<<

sustainability

Review Current Trends of Adsorption in Continuous Mode: Literature Review and Future Perspectives

Mariko A. Carneiro , Ariana M. A. Pintor * , Rui A. R. Boaventura and Cidália M. S. Botelho *

Laboratory of Separation and Reaction Engineering—Laboratory of Catalysis and Materials (LSRE-LCM), Department of Chemical Engineering, Faculty of Engineering, University of Porto, Rua Dr. Roberto Frias, 4200-465 Porto, Portugal; [email protected] (M.A.C.); [email protected] (R.A.R.B.) * Correspondence: [email protected] (A.M.A.P.); [email protected] (C.M.S.B.)

Abstract: Arsenic is a toxic element for humans and a major pollutant in drinking water. Natural and anthropogenic sources can release As into water bodies. The countries with the greatest arsenic contamination issues lack the affordable technology to attain the maximum permitted concentrations. Adsorption can be a highly efficient and low-cost option for advanced water treatment, and the development of new cheap adsorbents is essential to expand access to water with a safe concentration of arsenic. This paper aims to review the state of the art of arsenic adsorption from water in continuous mode and the latest progress in the regeneration and recovery of arsenic. The disposal of the exhausted bed is also discussed. Fixed-bed column tests conducted with novel adsorbents like binary metal oxides and biosorbents achieved the highest adsorption capacities of 28.95 mg/g and 74.8 mg/g, respectively. Iron-coated materials presented the best results compared to adsorbents under other treatments. High recovery rates of 99% and several cycles of bed regeneration were  achieved, which can aggregate economic value for the process. Overall, further pilot-scale research is  recommended to evaluate the feasibility of novel adsorbents for industrial purposes. Citation: Carneiro, M.A.; Pintor, A.M.A.; Boaventura, R.A.R.; Botelho, Keywords: arsenic; adsorption; fixed-bed column; regeneration; biosorbents C.M.S. Current Trends of Arsenic Adsorption in Continuous Mode: Literature Review and Future Perspectives. Sustainability 2021, 13, 1. Introduction 1186. https://doi.org/10.3390/ Arsenic (As) is a metalloid from the group of the , which can be su13031186 released into water bodies from both natural and anthropogenic sources. Arsenic appears in water mainly from a combination of natural processes, such as weathering reactions, Academic Editor: Danilo Spasiano biological activity, volcanic emissions, and fluvial transport of As-containing Received: 2 December 2020 deposits [1,2]. In addition to natural causes, anthropogenic activities such as the use of Accepted: 20 January 2021 Published: 23 January 2021 arsenic compounds in agriculture and livestock, as well as mining and industry, are strong causes of groundwater and surface water arsenic contamination [3]. The prevention of

Publisher’s Note: MDPI stays neutral arsenic water pollution from anthropological activities requires a proper water management with regard to jurisdictional claims in plan and actions from local authorities; however, controlling natural sources of As is more published maps and institutional affil- complex and relies upon mitigation technologies and strict management of prevention iations. measures [4]. Arsenic is a class 1 carcinogen, and human contamination can occur via consumption of contaminated water or food grown in contaminated soil. The most common inorganic forms are trivalent—As(III) and pentavalent—As(V), being the trivalent form much more toxic to human health [5]. The World Health Organization (WHO) and the European Copyright: © 2021 by the authors. µ Licensee MDPI, Basel, Switzerland. Directive 98/83/EC have established a guideline value of 10 g/L for As in drinking water. This article is an open access article Consumption of contaminated water with As is a major concern in developing coun- distributed under the terms and tries, which, in most cases, lack affordable technologies to attain WHO guidelines and EU conditions of the Creative Commons parametric value. Among various existing techniques, adsorption is the most inexpensive Attribution (CC BY) license (https:// and efficient process for advanced water and wastewater treatment [6]. Commercially creativecommons.org/licenses/by/ available adsorbents, such as activated carbon, present some disadvantages, such as high 4.0/). production cost and difficulty in regeneration, leading to an increase in treatment costs [7].

Sustainability 2021, 13, 1186. https://doi.org/10.3390/su13031186 https://www.mdpi.com/journal/sustainability Sustainability 2021, 13, x FOR PEER REVIEW 2 of 24

production cost and difficulty in regeneration, leading to an increase in treatment costs Sustainability 2021, 13, 1186 2 of 25 [7]. Worldwide studies are being conducted to develop more efficient, economic, and sus- tainable adsorbents to remove arsenic from water, with biosorbents being one of the most promising alternatives to remove toxic metals and metalloids from contaminated waters. Worldwide studies are being conducted to develop more efficient, economic, and sustain- Coating a common and cheap material to turn it into an adsorbent is a common prac- able adsorbents to remove arsenic from water, with biosorbents being one of the most ticepromising among researchers alternatives toto remove improve toxic adsorption metals and capacity metalloids and from increase contaminated adsorption waters. rate. The most commonCoating acoating common materials and cheap are material iron, alumina, to turn it and into other an adsorbent metals, is such a common as zirconium andpractice manganese. among Beyond researchers the to increase improve in adsorption adsorbent capacity capacity, and increaseiron-impregnated adsorption rate.adsorbents areThe non-toxic, most common low cost, coating and materials accessible are iron, in large alumina, quantities and other [8]. metals, such as zirconium andSome manganese. review Beyond articles the [8–11] increase have in adsorbentgathered capacity, the research iron-impregnated into arsenic adsorbents adsorption, but are non-toxic, low cost, and accessible in large quantities [8]. most of the studies discussed were conducted in batch mode, where the influent is treated Some review articles [8–11] have gathered the research into arsenic adsorption, but most sequentiallyof the studies (in discussedcycles) in werethe reactor conducted for a in pr batche-determined mode, where time the [12]. influent In continuous is treated mode, thesequentially influent and (in cycles)effluent in are the reactorcontinuously for a pre-determined pumped into time and [12 ].out In continuousof the reactor. mode, Figure 1 illustratesthe influent batch and and effluent continuous are continuously operation pumped modes. into and out of the reactor. Figure1 illustrates batch and continuous operation modes.

Figure 1. Illustration of batch and continuous treatment. Figure 1. Illustration of batch and continuous treatment. Adsorption in continuous mode is preferable for industrial and commercial pur- posesAdsorption [6], hence in the continuous importance mode of investigating is preferable the state for industrial of the art of and arsenic commercial adsorption purposes [6],studies hence conductedthe importance in continuous of investigating mode, most the commonly state of inthe a fixed-bedart of arsenic column. adsorption studies This review article aims to investigate the tendency of arsenic removal on continuous conducted in continuous mode, most commonly in a fixed-bed column. mode, based on an analysis of published articles in the last 20 years, and also the progress inThis regeneration review ofarticle exhausted aims bedto investigate and recovery the of tendency arsenic. of arsenic removal on continuous mode, based on an analysis of published articles in the last 20 years, and also the progress in regeneration2. Methodology of exhausted bed and recovery of arsenic. In this article, a scoping literature review was carried out to gather studies from the 2. Methodologylast 20 years about arsenic removal from water in continuous adsorption processes. A total of 145 articles were gathered, with 33% being from the last 5 years (2016–2020), 15% from 2011–2015,In this article, 38% from a scoping 2006–2010, literature and 14% review from 2000–2005. was carried The main out databaseto gather used studies in this from the lastresearch 20 years was about ScienceDirect, arsenic removal followed from by Springer water and in continuous Google Scholar. adsorption EndNote X9processes. was the A total of 145software articles used were to manage gathered, references. with 33% being from the last 5 years (2016–2020), 15% from 2011–2015,The majority38% from of 2006–2010, selected articles and are14% related from to2000–2005. experimental The studies main database of adsorption used in this researchin continuous was ScienceDirect, mode to remove followed arsenic by from Springer water. Aand few Google other articles Scholar. were EndNote selected X9 was theto software provide complementaryused to manage information. references. The studies were grouped by adsorbent type: metal (hydr)oxides (Section 3.1), zero-valent iron (ZVI) (Section 3.2), (Section 3.3), soilThe and majority rock (Section of selected 3.4), carbon-based articles are adsorbents related (Sectionto experimental 3.5), biosorbents studies (Section of adsorption 3.6), in continuousindustrial wastemode and to remove by-products arsenic (Section from 3.7 ),water. nanocomposites A few other (Section articles 3.8), were granular selected ferric to pro- videhydroxide complementary (GFH) and information. commercial adsorbents The studies (Section were 3.9), grouped and layered by double adsorbent hydroxides type: metal (hydr)oxides(LDH) (Section (Section 3.10). Figure3.1), zero-valent2 illustrates theiron types (ZVI) of adsorbents(Section 3.2), cited minerals in this article. (Section 3.3), soil and rock (Section 3.4), carbon-based adsorbents (Section 3.5), biosorbents (Section 3.6), industrial waste and by-products (Section 3.7), nanocomposites (Section 3.8), granular fer- ric hydroxide (GFH) and commercial adsorbents (Section 3.9), and layered double hydrox- ides (LDH) (Section 3.10). Figure 2 illustrates the types of adsorbents cited in this article. Sustainability 2021, 13, 1186 3 of 25 Sustainability 2021, 13, x FOR PEER REVIEW 3 of 24

FigureFigure 2. Diagram 2. Diagram of adsorbents of adsorbents types in As types continuous in As continuous adsorption studies. adsorption studies.

FromFrom the selected the selected literature, literature, information information about the recovery about the of arsenic recovery and ofthe arsenic regen- and the regen- erationeration of exhausted of exhausted beds bedsis summarized is summarized in Section in Section4. Similarly,4. Similarly, information information about the about the fate fateof of contaminated contaminated beds beds is issummarized summarized in Section in Section 5. 5.

3. Adsorption3. Adsorption of Arsenic of Arsenic in Continuous in Continuous Mode: Adsorbents Mode: Adsorbents Type Type 3.1.3.1. Metal Metal (hydr)oxides (hydr)oxides 3+ IronIron oxides oxides are created are createdby the precipitation by the precipitation of Fe , and the of resulting Fe3+, andsubstance the resultingchar- substance acteristics such as chemical structure, composition, and physical aspects can vary with differentcharacteristics preparative such techniques as chemical [13]. Iron structure, oxides and composition, hydroxides andare the physical most common aspects can vary with coatingdifferent substances preparative in sand columns techniques for arsenic [13]. removal Iron oxides [14–20] and due hydroxides to their affinity are with the most common arseniccoating and relatively substances higher in sandabundance columns in nature. for arsenic removal [14–20] due to their affinity with arsenicThe combination and relatively of iron higher with another abundance metal oxide in nature. is commonly investigated in col- umn operationThe combination mode to increase of iron arsenic with anotherremoval [12,21]. metal oxideIron-zirconium is commonly binary investigated oxide- in column coatedoperation sand (IZBOCS) mode to achieved increase a maximum arsenic removal adsorption [12 capacity,21]. Iron-zirconium of 28.95 mg/g by binary the oxide-coated Thomas model [15], while zirconium oxide-coated sand only achieved an adsorption ca- sand (IZBOCS) achieved a maximum adsorption capacity of 28.95 mg/g by the Thomas pacity of 0.042 mg/g using the same adsorption model [22]. modelDynamic [15 tests], while with zirconiumother metal-based oxide-coated adsorbents sand also show only good achieved results. an Zirconium adsorption capacity of metal–organic0.042 mg/g framework using the column same experiments adsorption were model able [ 22to reduce]. arsenic initial concen- tration fromDynamic 100 µg/L tests to 10 with µg/L, other for 2270 metal-based and 1775 bed adsorbents volumes (BV) also for As(III) show and good As(V), results. Zirconium respectivelymetal–organic [23]. Manganese framework performance column experimentswas also investigated wereable in column to reduce experiments arsenic initial concen- conductedtration fromwith natural 100 µg/L manganese to 10 µg/L, oxides for [24] 2270 and and manganese-coated 1775 bed volumes sand (BV)(MCS) for [14] As(III) and As(V), forrespectively arsenic removal. [23 ]. Manganese performance was also investigated in column experiments Table 1 shows characteristics of dynamic adsorption studies conducted with metal conducted with natural manganese oxides [24] and manganese-coated sand (MCS) [14] for (hydr)oxides. arsenic removal. Table1 shows characteristics of dynamic adsorption studies conducted with metal (hydr)oxides. Sustainability 2021, 13, 1186 4 of 25

Table 1. Comparison between dynamic tests using metal (hydr)oxides and zero-valent iron (ZVI) as adsorbents for arsenic removal.

Influent BV to Arsenic Max. Adsorption Breakthrough Adsorbent Concentration Breakthrough Ref. Species Capacity (mg/g) Time at 10 µg/L (h) (µg/L) Point of 10 µg/L Manganese-coated As 1000 0.079 250 pore volumes 18 days 1 [14] sand (MCS) Iron-Zirconium Binary Oxide-Coated As(V) 125 25.09–28.95 0.5–0.75 [15] Sand (IZBOCS) Natural iron mineral-quartz As(V) 500 7000 pore volumes [16] sand columns Crystalline hydrous As 320–400 14,000 [20] ferric oxide (CHFO) Zirconium As(III) 500 0.03310–0.04223 5.33–8.33 [22] oxide-coated sand

Zirconium metal As(III) 100 2270 [23] organic frameworks (UiO-66) As(V) 100 1775 [23] Hydrated stannic As(III) 1000 2400 [25] oxide (HSO) As(V) 1000 450 [25] As(V) 100 1900 [26] Zero valent iron (ZVI) 500–12,600 PV 2 [27] 20 [28] 1 Pilot-scale 2 For a breakthrough point of 1 µg/L.

The hybrid adsorbent of iron and zirconium (IZBOCS) achieved the highest value of adsorption capacity of 28.95 mg/g. The crystalline hydrous ferric oxide (CHFO) was able to treat a great amount of arsenic-contaminated solution (14,000 BV) with an outlet concen- tration below the WHO recommended value. The longer time of breakthrough of 18 days was achieved by manganese-coated sand (MCS) in a pilot-scale plant. Metal (hydr)oxides are usually impregnated in other materials to increase arsenic removal efficiency or/and reduce the treatment costs, as will be discussed in the following sections.

3.2. Zero-Valent Iron (ZVI) The mechanism and behavior of arsenic adsorption by zero-valent iron (ZVI) in dynamic systems was investigated in some studies [26–31]. Earlier investigations in column mode with zero-valent iron filings, which were carried out with an inlet arsenic solution of 100 µg/L (pH 7.0), led to an outlet concentration below 1 µg/L for 500 and 12,600 pore volumes (PV), with a time residence of 20 s and 25 min, respectively [27]. Results of dynamic adsorption studies conducted with zero-valent iron are shown in Table1. ZVI presented a performance similar to metal (hydr)oxides, however, the column studies with ZVI showed a substantial iron leaching of 8 to 17 mg/L in the final effluent [28,29], which can be a hindrance in the application of ZVI to remediate contaminated waters. The corrosion of iron by acid waters increases the iron release, which can result in an effluent with an iron concentration of 90–130 mg/L [28].

3.3. Minerals Column experiments to investigate arsenic removal were carried out using a variety of minerals as adsorbents: akaganeite [32], attapulgite [33], bauxite [34–37], goethite [38], hematite [16,39], mackinawite [40], magnetite [41–44], siderite [16,39,45], schwertman- Sustainability 2021, 13, 1186 5 of 25

nite [46], and zeolite [47,48]. Iron and aluminum-based minerals are considered good adsorbents due to their known affinity for heavy metals and metalloids and due to their abundance in nature. Iron modified calcined bauxite (MCB) was investigated in column experiments for arsenic removal from aqueous solution [34–37]. The breakthrough point of 10 µg/L for a 10 cm bed depth (2 cm diameter) was reached after 31 h (~270 BV) at 5 mL/min with an As(V) solution of 2 mg/L [34]. The breakthrough point (10 µg/L) for an As(III) solution of 1 mg/L, in similar conditions of flow rate and bed depth, was reached after 50 h (~470 BV) [36]. The MCB maximum adsorption capacities reported in column mode were 0.606 mg/g and 0.490 mg/g for As(V) and As(III), respectively [34,36]. A novel adsorbent consisting of magnetite-rich particles (MEP) was developed with a low magnetic intensity separation from mill scale, which is constituted of wustite (FeO), hematite (α-Fe2O3), magnetite (Fe3O4), elemental iron, and residual oil and grease [43]. Column adsorption tests were carried out in four sets of packed-bed columns, each with 3.5 cm diameter, 20 cm bed height and 550 g of MEP. A synthetic wastewater containing 500 µg/L of As(V) was fed into the column at different flow rates. Breakthrough point (~10 µg/L) was reached after about 17 d and 30 d, at a feed flow rate of 6.4 mL/min and 3.2 mL/min, respectively. After saturation, at the 48th d of operation, MEP columns were regenerated with 0.2 N NaOH for 24 h at a flow rate of 1 mL/min, and the sorption capacity was reinstated to the initial condition. Column tests with naturally occurring siderite, an iron carbonated mineral, led to a final effluent below 1.0 µg/L (26,000 pore volumes) using a solution containing 250 µg/L As(III) and 250 µg/L As(V) [45]. A similar study with a column packed with half siderite and half hematite content achieved As concentration of 1.0 µg/L after 7200 pore volumes (PV) [45]. The addition of hematite in siderite filters reduces the arsenic removal efficiency because siderite is usually more efficient to uptake As from water than hematite [49]. Siderite-hematite filters were activated with H2O2 which increased arsenic adsorption efficiency up to a throughput of 500 PV [39]. Three types of granular schwertmannite, an Fe(III)-oxyhydroxy sulfate mineral, were in- vestigated in column tests for arsenate removal [46]. The experiments were conducted in a lab-scale column with 2 cm diameter and 20 cm bed height, fed with simulated groundwater containing 210 µg/L of As(V). The irregular shape (IS) type achieved the best performance and breakthrough of 10 µg/L occurred after 8100 BV, followed by the cylindrical shape (CS) type with 4200 BV and the spherical shape (SS) type with 120 BV. The adsorption capacities of IS, CS, and SS were 0.93, 0.66, and 0.33 mg/g, respectively. Table2 shows the results of dynamic adsorption studies conducted with minerals. It is possible to highlight the performance of magnetite derived from mill scale compared to other column studies using minerals such as arsenic adsorbent. Natural siderite also achieved good results, even better than the siderite-hematite media. The other minerals, most of them modified iron, presented relatively similar results for arsenic removal.

Table 2. Comparison of dynamic tests using minerals as adsorbents for arsenic removal.

Influent BV to Arsenic Max. Adsorption Breakthrough Adsorbent Concentration Breakthrough Ref. Species Capacity (mg/g) Time at 10 µg/L (h) (µg/L) Point of 10 µg/L Iron modified calcined As(V) 2000–8000 0.470–0.606 269.2–300.64 2.15–59 [34] bauxite (MCB) Iron modified calcined As(III) 1000 428.02–489.17 28–96 [35] bauxite (MCB) Iron modified calcined As(III) 1000 0.392–0.459 427.85–489.17 28–96 [36] bauxite (MCB) Iron-modified As(V) 100 300 [47] clinoptilolite Sustainability 2021, 13, 1186 6 of 25

Table 2. Cont.

Influent BV to Arsenic Max. Adsorption Breakthrough Adsorbent Concentration Breakthrough Ref. Species Capacity (mg/g) Time at 10 µg/L (h) (µg/L) Point of 10 µg/L Mill-scale derived As(V) 100 3.60–5.00 >100 [42] magnetite particles Magnetite-enriched As(V) 500 17–30 days [43] particles (MEP) Mixed-valent iron As(V) 100–1000 10 888–1032 [44] oxide/magnetite As(III) + Natural siderite 250 + 250 1.090–2.000 11,600–26,000 PV [45] As(V) As(III) + Siderite-hematite 250 + 250 7200 PV [45] As(V) Activated Total As 500 0.177–0.185 8160 PV [39] siderite-hematite Goethite- polyacrylamide As 300 400 1 [38] composite (goethite-P(AAm) Granular As(V) 200 0.33–0.93 120–8100 [46] schwertmannite Iron-coated As(V) 2000 0.69 300 [48] zeolite (ICZ) Iron impregnated As(V) 397 charred granulated 200 [33] As(III) 175 attapulgite 1 For breakthrough point of 25 µg/L.

3.4. Soil and Rock Soils and rocks are also natural abundant materials, often containing minerals that enhance their arsenic uptake capacity. They stand out from other adsorbents due to their advantages of being low-cost and locally available. The arsenic uptake capacity of natural occurring laterite soil was studied in column operation mode using arsenic solutions and real samples [50–53]. Fixed-bed column studies conducted in a 2 cm diameter column and a bed depth of 30 cm, fed with arsenic solution of 500 µg/L, at 7.75 mL/min flow rate, reached the breakthrough point at 10 µg/L after 32 h for As (V) [52] and 18.5 h for As (III) [51]. The total arsenic removal by laterite soil, in similar column experiments (10 cm bed depth), with a groundwater (tube-well) sample of 0.33 ppm of arsenic was 98% [50]. Moreover, recent research has developed a novel laterite soil filter for arsenic that provides potable water to a population higher than 5000 people [54]. A model to predict the lifetime of the filter was validated by lab-scale and field-scale experiments. It was predicted that the novel laterite filter could last from 10 to 100 months, depending on flow rate, size, and the arsenic concentration of the feed water. The arsenic adsorption behavior of another soil, originated from limestone and heavily iron-coated, was investigated [55] in a glass column with 1 cm diameter, packed with 2 g of adsorbent (1.7 cm depth). The system was fed with an arsenate solution of 1 mg/L, in order 3− to investigate the effect of pH, pore water velocity, and the presence of (PO4 ). The phosphate produced a major effect on As(V) mobility and recovery, followed by the pore water velocity effect, with the least effect being observed due to pH variation. A few studies reported rocks as adsorbents for arsenic removal from water. Pisolite, a mining waste composed of iron and manganese aggregate, was applied for this pur- Sustainability 2021, 13, 1186 7 of 25

pose [56]. After 180 min of column operation, 1.0 g of pisolite removed 1.41 mg of As, and 1.0 g of activated (400 ◦C heating) pisolite removed 3.51 mg of As. Native limestones were investigated in column experiments fed by solutions with similar characteristics to As-rich well water from Mexico [57]. The authors observed the interference of bicarbonate and sulfate in As removal and indicated native limestones for the treatment of arsenic- contaminated water. Diatomite coated with hydrous ferric oxide (HFO) was able to treat about 1100 BV of As(III) solution (500 µg/L) until a breakthrough point of 10 µg/L [58]. The arsenic removal efficiency by an iron-oxide-coated natural rock (IOCNR) was studied in a fixed-bed column with 2 cm diameter, fed at an 8 mL/min flow rate at a pH of 5.7 [59]. The rock used in the study was Fe(III)-loaded granite. The breakthrough point of 10 µg/L for an initial As(III) concentration of 600 µg/L was reached after 31 h, 49 h, and 63 h for bed depths of 10 cm (~48 g), 15 cm (~73 g), and 20 cm (~97 g), respectively. A similar dynamic study was conducted with IOCNR, and the adsorption-exhausted sludge of As(III) and As(V) was mixed with Portland cement to reduce the impact and the cost of the water treatment process [60]. The study evaluated the As mobility in the material, and the leaching test allowed the authors to conclude that the adsorption column sludge mixed with Portland cement is an environmentally friendly material that could be applied in building construction. Table3 shows characteristics of dynamic adsorption studies conducted with soil and rock.

Table 3. Comparison of dynamic tests using soil and rock as adsorbents for arsenic removal.

Influent Breakthrough Max. BV to Arsenic Adsorbent Treatment Concentration Time at Adsorption Breakthrough Ref. Species (µg/L) 10 µg/L (h) Capacity (mg/g) Point of 10 µg/L Laterite soil None As 330 6.75 [50] Laterite soil None As(III) 500–1000 2.8–32 [51] Laterite soil None As(V) 500–1000 1.65–18.5 0.047 [52] Laterite soil None As(V) 500–1002 4.95–8.12 0.112–0.147 [53] Laterite soil None As 80–100 [54] Iron-oxide None As(V) 1000 15–875 [55] containing soil Thermic Pisolite As(V) 50,000 1.41–3.51 [56] (400 ◦C) Limestones None As(V) 1190–1340 3–15 weeks [57] Iron- incorporated As(V) 500 1100 diatomite Fe(NO ) ·9H O [58] 3 3 2 As(III) 500 1100 [Fe(25%)- diatomite)]

Iron-oxide- As(III) 600 31–63 [59] coated natural Fe(NO3)3·9H2O As(III) 300–1000 20–65 [60] rock (IOCNR) As(V) 1000–3000 5–18 Iron- impregnated FeCl As(V) 1000 16 [61] tablet ceramic 3 (ITCA)

Rock-based adsorbents presented better results for arsenic removal than soil-based ones. The longest breakthrough time was achieved by limestones, which maintained arsenic effluent solution under 10 µg/L for weeks, followed by IOCNR, which kept effluent concentration under the WHO standard for more than 60 h. Sustainability 2021, 13, 1186 8 of 25

3.5. Carbon-Based Adsorbents Most of the carbon-based adsorbent investigations in column operation mode for arsenic adsorption have been conducted with activated carbon (AC) or granular activated carbon (GAC) loaded with iron to benefit from the properties of both materials [62–67]. The use of AC, especially GAC, as a support for metal oxides/hydroxides, provides a more stable structure for As removal in comparison with the application of the metal oxide/hydroxide alone, mainly in fixed-bed column processes [63]. A continuous mode study using iron-impregnated GAC (GAC-Fe) and pristine GAC was conducted for As(V) removal [63]. The maximum As adsorption capacity observed was 0.470 mg/g for the GAC-Fe column, which performed approximately four times better than the GAC column of 0.133 mg/g [63]. Hydrous ferric oxide was immobilized onto GAC using phenol formaldehyde (PF) to provide an economically feasible solution for As removal from water [66]. Another methodology to prepare ferric AC was by precipitating amorphous HFO onto AC (FeO/AC) [65]. Laboratory column experiments with FeO/AC were performed at 8.33 mL/min with an As(V) initial concentration of 500 µg/L. FeO/AC was able to remove at least 99.8% of arsenic from the influent, and the breakthrough point of 50 µg/L was achieved after approximately four times more throughput volume compared to AC without iron loading. Table4 presents the results of dynamic adsorption studies using carbon-based adsorbents. The iron impregnation of GAC and AC was able to increase As uptake from water by about four times, considering results of column experiments [63,65]. However, GAC/AC are usually more expensive adsorbents, and their adsorption capacities obtained in dynamic experiments were found to be similar or inferior to other low-cost adsorbents such as minerals, soil and rock, biosorbents, and industrial waste and by-products.

Table 4. Comparison of dynamic tests using carbon-based adsorbents for arsenic removal.

Influent Max. BV to Breakthrough Arsenic Adsorbent Treatment Concentration Adsorption Breakthrough Time at Ref. Species (µg/L) Capacity (mg/g) Point of 10 µg/L 10 µg/L (h) Iron-modified Fe(NO ) ·9H O As 40–60 5600–34,000 [62] activated carbon 3 3 2 Iron-impregnated granular activated FeCl2 As 100–500 0.305–0.470 313–1494 [63] carbon (GAC-Fe) Granular activated As 100 0.118–0.133 380–388 [63] carbon (GAC) Granular activated carbon FeCl As(V) 180 2.873 [64] doped with iron 3 (Fe/GAC) Hydrous ferric oxide incorporated onto GAC with phenol FeCl /PF As(V) 1886 0.7117 180 1753 [66] formaldehyde 3 resins coating (HFO-PF- coated GAC) Iron-containing As(V) 57.2 7500 granular FeCl [67] 2 As(III) 56.1 7500 activated carbon Sustainability 2021, 13, 1186 9 of 25

3.6. Biosorbents Biosorbents are naturally occurring materials or waste biomass, and their high ad- sorption capacity and cost-effectiveness make biosorption a promising process in water and wastewater treatment [68]. The group of biomaterials includes aquatic biopolymers, plant biomass, microbial biomass, biochars, and agricultural waste [69]. Column stud- ies have investigated the potential of these materials, with or without pre-treatment or modification, to remove arsenic from aqueous solution in continuous mode [7,70–97].

3.6.1. Aquatic Biomaterials Chitosan Chitosan (CS) is a natural aminopolysaccharide originating from exoskeletons of arthropods [7] such as crabs and shrimps. The biopolymer is obtained by N-deacetylation of chitin [78] and has been applied in heavy metals and arsenic uptake due to its free amino groups that can serve as sites of coordination binding [72]. Chitosan also presents the advantage of being low cost, biodegradable, renewable, and non-toxic [75]. Early experimental column studies with chitosan beads investigated the feasibility of using chitosan without any pretreatment to remove arsenic from wastewater discharged from gallium arsenide (GaAs) manufacturing in continuous mode [72]. Laboratory tests were developed in a glass column with 1.8 cm diameter and 12 cm height with a total pack- ing volume of 30.5 mL, packed with 20 g of chitosan beads of 2.5 mm size. The industrial wastewater contained 2.5 mg/L of As(III) and 35 mg/L of As(V) and presented a pH of 4.8. Comparison tests were made with synthetic wastewater containing 10 mg/L of As(III) or As(V), at pH 5. Recycled chitosan beads were also investigated in continuous mode and presented similar breakthrough time to native chitosan beads. The experiments made with synthetic wastewater at 8 mL/min resulted in an adsorption capacity of 1.78 and 1.90 mg As/g bead, for As(III) and As(V), respectively. According to this study, chitosan is a promising adsorbent for arsenic removal from water. Later, molybdate-impregnated chitosan beads (MICB) were developed to improve arsenic adsorption capacity [73]. Laboratory column tests were carried out with synthetic effluent and GaAs manufacturing wastewater with similar characteristics to the previous study [72]. The breakthrough time at 30 mL/min was approximately 7 h for the synthetic wastewater and between 5 h and 8 h for the industrial wastewater. Recycling tests were made, and the adsorption curve remained similar up to 10 cycles of sorption–desorption. MICB adsorption capacity observed with synthetic wastewater was 1.98 and 2.00 mg As/g MICB for As(III) and As(V), respectively. The adsorption capacities observed for MICB were only slightly higher than the ones observed for chitosan without pretreatment, so the costs of adding molybdate were not justified by an incrementation in the adsorption capacity. In addition to molybdate impregnation for chitosan beads, studies have proposed other metal or inorganic oxide treatments to improve their arsenic adsorption capacity. Iron oxides are the most common [7,74–76,78,83], followed by aluminum oxides [70,74], magnesium oxide [81], and the combination of iron and manganese oxides [7]. Among the column studies of chitosan treated with metal oxides, the best arsenic adsorption capacity was observed for chitosan entrapped with Fe-Mn water treatment residual (C-WTR), with an As(III) removal capacity of 32.64 mg/g of C-WTR even after six cycles of sorption/desorption with 0.5 NaOH solution [78]. In addition to the treatment of chitosan with metal oxides, a simple method of proto- nation of amine groups, improved chitosan arsenic adsorption capacity on column tests to a range of 36–50 mg As/g of chitosan [71,80], which is the highest adsorption capacities among other types of chitosan modifications reported in column studies.

Alginate Alginate is a biopolymer composed of α-l-guluronate and β-d-mannuronate and is a major constituent of the cell wall of seaweed and sea tangle [77,98]. Alginate acids have high affinity with divalent cations and produce thermally irreversible and hydrophobic Sustainability 2021, 13, 1186 10 of 25

gels [84,99]. Alginate beads are applied to heavy metals and metalloids removal, and their preparation is usually made with calcium as cation [84]. Arsenic adsorption column studies using alginate treated with metal oxides, such as aluminum oxide [77] and iron oxides [82,84,100] have been reported. An alginate-based adsorbent containing calcinated alum sludge from a water treat- ment facility was developed to enhance surface area and overcome alginate bead limitation of slow adsorption kinetics [77]. The new adsorbent performed better than a commercial adsorbent (Bayoxide E33) by reaching 10 µg/L after 200 bed volumes, 50% higher than Bayoxide E33. The initial solution concentration was 200 µg/L. Fe(III) crosslinked alginate nanoparticles were also investigated in fixed-bed column experiments for As(V) removal [101]. The effect of inlet arsenate concentration (0.5, 1.0, and 1.5 mg/L), bed height (2, 4, and 6 cm) and flow rate (0.25, 0.50, and 1.0 mL/min) were studied. The Fe-alginate maximum adsorption capacity of 0.066 mg/g was achieved with the following dynamic experimental conditions: inlet As(V) solution of 0.5 mg/L, bed height of 2 cm, and flow rate of 0.50 mL/min.

Fish Scale The atlantic cod fish scale has been reported as an alternative biosorbent to remove heavy metals and arsenic from water [79,102,103]. Column tests were conducted using a dried fish scale in a glass column with 2.54 cm diameter and 23.88 cm bed height. The column was fed with a solution of 520 µg/L As(III) or 342 µg/L As(V) at a flow rate of 2.0 mL/min. The breakthrough point (C/Co = 0.1) was reached after 21.75 h and 25 h for As(III) and As(V), respectively.

3.6.2. Agricultural Waste Agricultural waste materials are mainly composed of lignin and cellulose, combined with other components such as hemicellulose, extractives, lipids, proteins, starches, water, hydrocarbons, and others [104]. This type of waste is a potential source of adsorbents for heavy metals and metalloids. The majority of research is directed to cationic heavy metals, and few studies have investigated the capacity of various food and agricultural wastes for the biosorption of arsenic [9]. Carbonized sugarcane bagasse (SCC) is an alternative to active char carbon prepared from agricultural waste material [94]. Column tests were performed to investigate the arsenate and arsenite removal capacities of SCC modified by thioglycolic acid impregna- tion [94]. Breakthrough curves were generated at different bed heights (11.3 cm to 32.1 cm), flowrates (3.0 to 7.0 mL/min), and arsenic initial concentration (500 to 1500 µg/L) with a pH of 6.0. Experimental and predicted (by Thomas model) adsorption capacities for As(III) and As(V) were found. Maximum adsorption capacities of 0.085 mg/g and 0.084 mg/g were reached for As(III) and As(V), respectively, at a flow rate of 3.0 mL/min, bed height of 32.1 cm and initial concentration of 1500 µg/L.

Rice-Based Adsorbents Rice polish, a waste from rice milling industries, was investigated as a potential adsorbent in up-flow fixed-bed column studies [93]. The effect of design parameters, such as bed height (5 to 25 cm), flow rate (1.66 to 8.33 mL/min), and initial metal concentration (100 to 1000 µg/L), was studied. Maximum uptakes of 0.067 mg/g for As(III) and 0.079 mg/g for As(V) were achieved with 25 cm bed height and 30.5 g of rice polish, at 1.66 mL/min flow rate, fed with a 1000 µg/L initial arsenic concentration. Desorption–sorption cycles were also investigated, and the recovery efficiency achieved values of approximately 99%. A good adsorption of arsenic performance was observed for rice waste studies in continuous mode [93,95]. Considering that rice production is strong in some places where arsenic contamination is a major problem, such as Bangladesh and other Asian countries, Sustainability 2021, 13, 1186 11 of 25

rice by-products can be a low-cost potential solution for arsenic removal from water in these regions.

Cotton-Based Adsorbents Column studies with cotton-based adsorbents were performed after iron modification of the materials [91,97]. A novel adsorbent, Fe(III)-loaded ligand exchange cotton cellulose [Fe(III)LECCA] was synthetized for selective arsenate adsorption [97]. In column tests with 1 mg/L arsenic aqueous solution, at pH of 7.10, a breakthrough capacity (at 10 µg/L) of 5.3 mg/g was achieved with a flow rate of 26 BV/h. In similar column tests with tap water spiked with As(V), the breakthrough point adsorption capacity was reduced to 65% of the original value. Fixed-bed column experiments were also performed with another novel adsorbent, consisting of Fe-based metal-organic framework coated onto cotton fiber composites [MIL-88A(Fe)/cotton fibers] [91]. A solution with an As(V) concentration of 500 µg/L was fed at 6.0 mL/min flow rate to the column with 34 cm length and 1.6 cm diameter. The breakthrough point of 10 µg/L was achieved after approximately 440 bed volumes. Another cotton-based adsorbent study in column mode was reported, but instead of iron, zirconium was loaded to the material to improve uptake capacity [87]. The novel adsorbent was synthesized by radiation-induced graft polymerization (RIGP) of a phos- phoric monomer onto nonwoven cotton fabric. Zr(IV) was then loaded to the phosphoric units packed in the column. The new Zr-type adsorbent arsenic removal capacity was investigated in column tests at different pH (1, 2, 3, 5, 7, and 9) and different flow rates (200 and 1000 h−1) with an As(V) solution of 1 mg/L. The maximum adsorption capacity of 0.749 mg/g was obtained at pH 2 and 1000 h−1 flow rate. Among the agricultural wastes used in dynamic studies for arsenic removal, the cotton- based materials led to better adsorption capacities because of modifications with iron and zirconium. Moreover, iron loading has provided higher adsorption capacity to cotton-based adsorbents than zirconium.

3.6.3. Biochar Biochar is a stable black carbon material produced from a wide range of biomass types [105]. The biochar characteristics such as large specific surface area, porous structure, enriched surface functional groups, and mineral components, combined with its being environmentally friendly, low-cost, and effective, make the application of biochar as an adsorbent attractive to researchers [105,106]. Recent studies reported the association of iron composites to biochar in order to improve arsenic adsorption removal in column mode [41,88,96]. Biochar derived from hickory chips was impregnated with iron nanoparticles by direct hydrolysis of iron salt and investigated for arsenic adsorption–desorption [88]. A column measuring 5 mm in diameter and 12 mm of bed height was filled with 1 g of the iron- impregnated biochar and fed with a 50 mg/L As(V) solution at 2 mL/min. Column experimental data were well fitted by the convection-dispersion-reaction (CDER) model. Desorption tests carried out with 0.05 mol/L NaHCO3 obtained an 85% removal of the retained arsenic. Magnetized biochar composite (MBC) was prepared by precipitating magnetite pri- mary nanoparticles (Fe3O4) on Douglas fir biochar [41]. Fixed-bed column experiments were performed using a 10 mg/L As (III) solution at pH 7 at 2 mL/min flow rate, and the breakthrough point was reached after 40 min. Regeneration tests were conducted with 0.5 M KH2PO4 at pH 4.5, in three cycles, and results show that restoration rates were 85%, 81%, and 76% from the initial adsorption capacity. The desorption efficiency was reduced from 87% to 80% and 76%. Recently, iron oxide nanoneedles array-decorated biochar fibers (Fe-NN/BFs) were prepared by a hydrothermal method from waste cotton fiber [96]. Fixed-bed columns containing 2 g of adsorbent were fed with 275 µg/L arsenic spiked natural groundwater Sustainability 2021, 13, 1186 12 of 25

at 2.26 mL/min and at pH 6. For a breakthrough point of 10 µg/L, the treated volumes for As(V) and As(III) were 1350 and 350 bed volumes, respectively. For two columns in tandem, the treated volume increased to 2900 BV for As(V) and 2500 BV for As(III). The adsorption capacity of the column was 1.80 and 1.55 mg/g for As(V) and As(III), respectively. Five cycles of regeneration (0.1 M NaOH) were conducted with effective treatment performance.

3.6.4. Fungal Biomass A recent study investigated the arsenic adsorption capacity of four novel fungal strains in column mode [89]. Westerdykella sp. (FNBR_3), Trichoderma sp. (FNBR_6), Lasiodiplodia sp. (FNBR_13), and Rhizopus sp. (FNBR_19) were isolated from agricultural soils containing arsenic in West Bengal, India. Column experiments were performed in a glass column (0.7 cm diameter and 15 cm bed height) packed with alginate beads containing 10 g of immobilized fungal biomass at pH 6 with a flow rate of 8.5 mL/min and arsenic initial concentration of 200 mg/L. Lasiodiplodia sp. (FNBR_13) led to the highest values of breakthrough time (Cb = 100 µg/L) and adsorption capacity: 110 min and 74.88 mg/g, respectively. The Thomas model was fitted to the experimental results and the highest correlation, 0.9799, was exhibited by FNBR_13 experimental data. In another study, Aspergillus niger fungal biomass was treated with iron oxide to remove As(III) and As(V) in continuous mode [92]. A mass of 6.925 g of iron oxide-coated biomass (IOCB) beads were packed in a column with 1.25 cm diameter and 23.7 cm effective bed depth. Different solutions of 100 µg/L As(III) or As(V) at pH 6 were fed to the column at 3.1 mL/min flow rate. The breakthrough point of 10 µg/L for As(V) was achieved after 800 bed volumes for IOCB and after 130 bed volumes for regenerated IOCB (0.25 M NaOH at 10 mL/min). The As(III) breakthrough point could not be achieved in the first cycle, but it was achieved with regenerated IOCB after 340 bed volumes. The adsorption capacities of IOCB were reported to be 1.08 mg/g for As(V) and 1.21 mg/g for As(III). Fungal biomass can be considered a good source of biosorbents for arsenic removal. However, high values of adsorption capacity reported in the first study [89] can be ex- plained by the high arsenic initial concentration of 200 mg/L, 20,000 times higher than the initial arsenic solution in the second study [92]. The second study reported that fungal biomass with iron modification was able to achieve adsorption capacities in continuous mode similar to other biosorbents’ removal capacity.

3.6.5. Plant Biomass The arsenic adsorption capacity of sorghum biomass (SB), one of the greatest cereal crops in the world, was investigated in continuous mode [86]. Two columns with 5.4 cm diameter and 40 cm active bed height were fed with a solution containing 5 mg/L of As(III) at 10 mL/min. One column was packed with 150 g of non-immobilized sorghum biomass (NISB) and the other one with 140 g of immobilized sorghum biomass (ISB). The maximum adsorption capacities observed for NISB and ISB were, respectively, 2.765 and 2.437 mg/g. Preliminary experiments were conducted with immobilized Garcinia cambogia biomass beads (IGCFIX) to remove As(III) in continuous-flow conditions [90]. A column with 1.6 cm diameter and 7 cm of packed bed height was fed with a 100 mg/L As(III) solution at a flow rate of 13.3 mL/min. A regeneration study was carried out with 0.05 M NaOH at 1 mL/min flow rate in five cycles. The results achieved point to the feasibility of using IGCFIX in column operation. Table5 shows characteristics of dynamic adsorption studies conducted with carbon- based biosorbents. The average performance of biosorbents in column studies was superior to that of traditional types of adsorbents, such as metal (hydr)oxides and ZVI (Table1 ) minerals (Table2), soils and rocks (Table3), and carbon-based materials (Table4). The dif- ferent categories of biosorbents showed similar results but more studies were conducted with aquatic biomaterials, which can justify their higher feasibility to column studies. Sustainability 2021, 13, 1186 13 of 25

Individually, the cod fish scale and chitosan-based materials showed higher maximum adsorption capacities than other biosorbents.

Table 5. Comparison of dynamic tests using biosorbents for arsenic removal.

Influent BV to Arsenic Breakthrough Max. Adsorption Adsorbent Concentration Breakthrough Ref. Species Time at 10 µg/L (h) Capacity (mg/g) (µg/L) Point of 10 µg/L Aquatic Biomaterials Chitosan As(V) 30,000–120,000 424.7–4444.6 36–50 [71] Chitosan As(V) 120,000 4.33 [80] As(V) 10,000 1.9 Chitosan [72] As(III) 10,000 1.78 Molybdate- As(V) 10,000 1.99 impregnated [73] As(III) 10,000 1.96 chitosan (MICB) Chitosan-based porous magnesia-impregnated As(V) 300 42–100 17.2 [81] alumina (MIPA) Porous Fe-chitosan As(III) 975 1.19 3000 [83] beads (P/Fe-CB) Iron chitosan spacer As(V) 500–1000 59.0–106.0 210 [76] granules (ICS) As(III) 500–1000 50.5–107.7 132 As(V) Waste Fe/Mn oxides 500 As(III) 13.84–13.97 into chitosan beads 500 2700 [78] As(III) + 30.02–32.64 (C-WTR) 500 + 500 As(V) Magnetic binary oxide particles (MBOP) As(III) 1000 26.15–72 0.436–0.477 [74] using chitosan Granular adsorbent (GA) using back-wash As(V) 150 210 [7] residue material and chitosan Iron oxide As(V) 230 50 0.0138 [84] loaded alginate As(III) 45 Ferric hydroxidemicrocapsule As(III) 100 75 days [82] loaded calcium alginate (FHMCA) Atlantic cod fish scale As(V) 342–520 21.75–25 4.87–38.36 [79] Agricultural Waste Iron(III)-loaded phosphorylated orange As(V) 15,800 81.4 [85] juice residue (POJR) Fe-based metal-organic framework/cotton As(III) 500 320–440 [91] fibers composites As(V) 3–30.83 0.025–0.079 Rice polish 100–10,000 [93] As(III) 2.66–29 0.030–0.067 Rice husk As(III) 50–500 0.016–0.022 [95] Sustainability 2021, 13, 1186 14 of 25

Table 5. Cont.

Influent BV to Arsenic Breakthrough Max. Adsorption Adsorbent Concentration Breakthrough Ref. Species Time at 10 µg/L (h) Capacity (mg/g) (µg/L) Point of 10 µg/L Fe(III)-loaded ligand exchange cotton As(V) 1000 5.3 ~400 [97] cellulose adsorbent [Fe(III)LECCA] Nonwoven cotton fabric As(V) 1000 0.749 270–1420 [87] Thioglycolated As(V) 1500 0.084 sugarcane [94] As(III) 1500 0.085 carbon (TSCC) Biochar Iron-impregnated As 50,000 0.167 [88] biochar Magnetite precipitated As(III) 10,000 2.85 [41] onto Douglas fir biochar Iron oxide nanoneedle As(V) 275 1.80 1350 array-decorated biochar [96] As(III) 275 1.55 350 fibers (Fe-NN/BFs) Fungal and Plant Biomass Non-immobilized sorghrum 5000 27 1 2.765 31 pore volumes 1 [86] biomass (NISB) Immobilized sorghum 5000 24 1 2.437 27 pore volumes 1 [86] biomass (ISB) Novel fungal strains + As 200,000 0.50–1.83 2 59.5–74.8 [89] alginate beads Iron oxide-coated As(V) 100 1.080 800 [92] fungal biomass (IOCB) As(III) 100 1.210 3 340 3 1 For breakthrough point of 50 µg/L 2 For breakthrough point of 100 µg/L 3 Regenerated bed.

3.7. Industrial Waste and By-Products Looking for industrial waste materials for arsenic uptake and other pollutants’ removal from water is of great importance to improve sustainability because it can be cost-effective and also reduce waste disposal in industry. Column experiments have shown good potential for arsenic adsorption using sugar industry waste [107], water treatment solid residuals [77,108], cement [109–111], mining residual products [57], and, more recently, waste and by-products from scrap metal recycling [112] and the steel industry [42–44]. Regarding sugar industry waste, column experiments for arsenic removal were devel- oped using, as adsorbent, bagasse fly ash (BFA) [107]. The column data were analyzed by bed depth service time (BDST) and Yoon and Nelson models. The arsenate and arsenite removal efficiencies in synthetic influents were 98.9% and 95.6% and for real arsenic-contaminated water were 91.23% and 87.76%, respectively. The authors concluded that the BFA adsorption column is applicable to natural water remediation as regards As contamination. Chemical sludge rich in iron from a water treatment plant (WTP) in the USA was also investigated as a potential adsorbent for arsenic [108]. The materials were rich in iron because the WTP used ferric sulfate as coagulant. Column tests were conducted using the ferric sludge and granular ferric hydroxide (GFH) for comparison purposes. The ferric residual presented a performance similar to GFH and achieved a breakthrough point of 10 µg/L after 26,400 bed volumes, for a groundwater sample containing 37.7 µg As/L. Sustainability 2021, 13, 1186 15 of 25

Table6 shows characteristics of dynamic adsorption studies using industrial waste and by-products as adsorbents.

Table 6. Comparison of dynamic tests using industrial waste and by-products as adsorbents for arsenic removal.

As Influent Max. As Remov. Breakth. BV to Breakth. Adsorbent Treatm. Species Concentration Adsorption (%) Time at Point of Ref. (µg/L) Capacity (mg/g) 10 µg/L (h) 10 µg/L Bagasse fly As(V) 50 98.9 [107] None ash (BFA) As(III) 50 95.6 [107] Water treatment residual solids None As 37.7 26,400 [108] Iron As(V) 2000–4000 3.8852 4.0–41.5 [109] oxide-coated Fe(NO3)3·9H2O cement (IOCC) As(III) 500–2700 0.1462–0.3082 54.04–84.77 [110] Hardened paste of Portland Cement + water As(V) 500 >90 [111] cement (HPPC)

Stainless None 10,000 84% steel slags As(V) [112]

IOCC presented an adsorption capacity similar to other traditional adsorbents such as carbon-based materials and metal (hydr)oxides. In terms of arsenic removal efficiency, the iron modified cement (IOCC) does not present advantages compared to the hardened paste of Portland cement (HPPC), even the studies being conducted with different arsenic species. Use of industrial waste and by-products can be an economic and sustainable way to remove arsenic from water, but further studies should be carried out on their viability for practical applications.

3.8. Nanocomposites Nano adsorbents are very promising and highly efficient materials for arsenic removal from groundwater; however, separation and regeneration are still challenges to their practical applica- tion [11]. Nanoparticles are <100 nm in size [113], and different nanomaterials have been tested for arsenic adsorption. Column studies have been conducted with nanoparticles of different ma- terials: alginate [101], aluminum [114], carbon [115–117], copper [118,119], iron [21,96,120–122], manganese [123], polymers [114,124,125], zirconium [126], and titanium [21,127,128]. In order to facilitate reuse and regeneration, the nanoparticles are usually coated onto another materials such as glass beads [127], gel beads [114,125], polystyrene [124], and alginate beads [82]. Concrete-based iron oxide (IO) nanoparticles (NP) were tested in laboratory exper- iments using a glass column of 2.2 cm in diameter and different bed heights (10, 20, and 30 cm) [122]. Contaminated water containing 10 mg/L of As(V) was fed constantly at 7.5 mL/min flow rate. The time of breakthrough for 10 µg/L of arsenate was found to be 4.1, 6.5, and 8.6 h for 10, 20, and 30 cm of bed height, respectively. Multi-walled carbon nanotubes (CNTs) were also investigated in column adsorption tests for As(III) and As(V) removal [115]. A glass column containing 2.0 g of CNT and 2.0 g of washed river sand was operated under different flow rates (20, 30, and 40 mL/min) at pH 6, with an initial arsenic concentration of 40 µg/L. The optimal flow rate for operation was found to be 30 mL/min leading to an adsorption capacity of 0.014 mg/g for As(V) and 0.0135 mg/g for As(III). The regeneration process was conducted with different eluents, and 1.0 M HCl achieved the best results. A more recent study was conducted with another form of carbon-based sorbent: multi-functionalized magnetic graphene [117]. Adsorption-desorption column tests for As(III) and As(V) were taken with different flow rates, using NaOH and HCl as eluents, and results showed an arsenic recovery of approximately 98%. Sustainability 2021, 13, 1186 16 of 25

Table7 presents the results of dynamic adsorption studies conducted with nanocom- posites. They are claimed to be a highly efficient option for metals and metalloids adsorp- tion; however, the results of tests developed in continuous mode do not prove that yet.

Table 7. Comparison of dynamic tests using nanocomposites, GFH, and commercial adsorbents for arsenic removal.

Influent BV to Arsenic Max. Adsorption Breakthrough Adsorbent Concentration Breakthrough Ref. Species Capacity (mg/g) Time at 10 µg/L (h) (µg/L) Point of 10 µg/L Nanocomposites Microwave-assisted economic multi-walled As(V) 0.014 40 50–100 [115] carbon nanotubes As(III) 0.0135 (MWCNTs) Iron doped phenolic resin-based activated As(V) 1000 2.78–3.24 [116] carbon micro-nano As(III) 1000–3000 2.60–3.30 particles Cupric oxide (CuO) As 109 7–15 [119] nanoparticles As(V) 120 α-MnO nanofibers 200 [123] 2 As(III) 200 Concrete-maghemite As(V) 10,000 4.1–8.6 [122] nanocomposites

Metal-doped titania As 250–1000 0.477–0.61 1.67–6.5 [127] nanoparticles coated glass beads As(III) 250 1–3.67 [127] Zirconium-organic frameworks@biomass- As(III) 500 1.667 500 derived porous [126] As(V) 500 2.167 651 graphitic nanocomposites (Ui) Granular ferric hydroxide (GFH) and commercial adsorbents As(V) 800 4.92–6.07 [129] GFH As 10,000 ~500 [130] As 500 More than 40,000 [130] Ferrosorp plus (FP) 0.79 728 1 [131] ArsenXnp—a hybrid As 85 29,000 2 [120] anion exchanger Adsorbsia GTO As 28 3000–10,000 [128] Iron nanoparticle resin As(V) 500 3.229 3512–4000 [132] (Lewatit FO36) Fe (III)-phosphorylated As 80 6 (for 3 µg/L) [133] resin (Fe-XAD8-DEHPA 1 For breakthrough point of 3.75 µg/L 2 For breakthrough point of 50 µg/L.

3.9. GFH and Commercial Adsorbents Granular ferric hydroxide (GFH) is a well-known, commercially available adsorbent used for arsenic and other water pollutants removal; however, it has some limitations for long-term application [134]. Column studies have been conducted to investigate specific characteristics of an arsenic adsorption process with GFH [13,129–131,135–140]. Rapid small-scale tests were performed using three commercially available adsorbents— E33, GFH, and Metsorb—to assess the interference of silica, phosphate, vanadate, and pH Sustainability 2021, 13, 1186 17 of 25

in each of them [137]. The GFH was found to be the most susceptible to ionic interference. In addition to that, the pH and silica had the greatest impact on arsenic adsorption com- pared to vanadium and phosphate. Another lab-scale study showed that organic matter reduced the adsorption capacity of GFH [129]. Real-time pilot-scale tests were conducted with GFH treating an inlet arsenic concentration of approximately 1000 µg/L at a flow rate of 25 mL/min, and X-ray fluorescence (XRF) was validated as a monitoring method for arsenic accumulation through the column [136]. Resins are also popular commercial adsorbents due to their excellent hydraulic proper- ties and mechanic resistance [133]. The combinations of resins with metal (hydr)oxides are usually known as hybrid systems. They gather the selectiveness of the metal with the char- acteristics of the resins to result in more efficient products with better adsorption capacity. A commercial hybrid anion exchanger named ArsenXnp is formed by the combina- tion of anion exchange resin beads with hydrated ferric oxide (HFO) nanoparticles [120]. Field columns packed with ArsenXnp could run more than 20,000 bed volumes before the breakthrough point of 50 µg/L of arsenic. The regeneration process after saturation was carried out using 2% NaCl and 2% NaOH and allowed multiple cycles of reuse. Table7 shows characteristics of dynamic adsorption studies conducted with GFH and commercial adsorbents. Particularly good results were achieved in terms of treated bed vol- umes, using some alternative materials, such as biosorbents and industrial waste. They can treat similar amounts of contaminated water, in a more sustainable and economic way.

3.10. Layered Double Hydroxides (LDH) Layered double hydroxides (LDH) are a class of anionic clays that have been employed more recently to remove such as As(V) and As(III) from aqueous solution [141]. Despite their good adsorption characteristics, the use of LDH in fine powder and its further separation after treatment is still a hindrance for their wide application [142]. There are various forms of LDH; however, few of them have been applied in dynamic tests for arsenic removal. A zirconium and iron layered double hydroxide (Zn-Fe-LDH) was prepared by a modified co-precipitation method, and its arsenic removal capacity was studied in column tests [143]. A column (2.0 cm diameter × 10 cm height) packed with 0.8 g of adsorbent was fed with As(V) solution of 1 mg/L, at a flow rate of 0.6 mL/min and pH 7. The handling capacity achieved for the breakthrough point of 10 µg/L was 12.5 L/g of adsorbent. Desorption tests showed that a solution of 0.5 M NaOH was able to remove 67.8% of As(V) from the adsorbent. Fixed-bed experiments were also conducted with calcinated Mg-Al-LDH to investigate the effect of particle size, pH, initial As(V) concentration, and flow rate in breakthrough curves [144]. The decrease in particle size, flow rate, and initial concentration all caused an increase in breakthrough time. Results of adsorption studies, conducted with LDH adsorbents in continuous mode, are presented in Table8. Although LDH are composed of metals with strong affinity to arsenic, the reported column studies did not show interesting results if compared with other metal-containing adsorbents discussed in this article.

Table 8. Comparison of dynamic tests using LDH as adsorbents for arsenic removal.

BV to Influent Breakthrough Time Adsorbent Arsenic Species Breakthrough Ref. Concentration (µg/L) at 10 µg/L (h) Point of 10 µg/L Zn-Fe-LDH As(V) 1000 ~300 [143] Mg-Al-Cl-LDH Total As 506 10 [142] Mg-Fe-Cl-LDH Total As 506 7 [142] Sustainability 2021, 13, 1186 18 of 25

4. Recovery and Regeneration The recovery of arsenic and regeneration of the adsorbent are relevant to ensure the viability of the adsorption process on a large scale. The balance between the cost of regeneration and the cost of new adsorbent should be taken into consideration when choosing adsorption for arsenic removal from contaminated waters. Some of the studies discussed in the previous sections presented the results of adsorption/desorption cycles in continuous mode to evaluate the increase in adsorbent life span. Most desorption experiments have been conducted using NaOH [7,20,33,42,47,50– 52,60,70,74,76,78,81,90,93,96,100,117,127,132,140] as eluent with different concentrations, showing that arsenic recovery is possible with a basic solution. However, some studies reported that alkaline treatment removed coated iron from the adsorbent, and consider- ing that most treatments are iron-based, the concentration of eluent should be adjusted depending on each adsorbent’s characteristics. Other chemicals were also tested in the recovery process: sodium bicarbonate (NaHCO3)[88], sodium fluoride (NaF) [80], sodium chloride (NaCl) [71,80], potassium hydroxide (KOH) [95], monopotassium phosphate (KH2PO4)[41], phosphoric acid (H3PO4)[72,73], hydrochloric acid (HCl) [72,73,85,117], and sulfuric acid (H2SO4)[72,73]. Some adsorbents were successfully regenerated after 320 cycles [73], and others after only two cycles [75,76]. Good arsenic recovery rates were observed in column stud- ies. For example, with soils and rocks, recovery ranged from 92% to 99% [51,52,55,59]. Some biosorbents and nanocomposites also presented higher arsenic recovery, such as rice polish (99.6%) [93] and cupric oxide nanoparticles (99.4%) [119].

5. Disposal of Contaminated Adsorbents The sustainable application of novel adsorbents on a real scale also relies upon the fate of the contaminated bed. The toxicity characteristic leaching procedure (TCLP) is commonly applied to evaluate whether the waste can be safely disposed. According to USEPA 1992, the disposal of arsenic waste can occur in domestic landfills if the leachate concentration of arsenic is less than 5.0 mg/L; otherwise, the waste should be disposed of in a hazardous waste landfill [74]. TCLP results reported with arsenic-contaminated beds are presented in Table9.

Table 9. TCLP results from arsenic-contaminated beds.

Maximum Arsenic Type of Adsorbent Adsorbent Under USEPA Limit? Ref. Leached (mg/L) Ferric hydroxide microcapsule-loaded Biosorbent 0.3 Yes [82] alginate beads (FHMCA) Biosorbent Iron doped chitosan spacer granules (ICS) 0.02 Yes [76] Magnetic binary oxide particles (MBOP) Biosorbent 0.07 Yes [74] (solidified and stabilized) Carbon-based Iron-containing granular activated carbon 0.09 Yes [67] Carbon-based HFO-PF-coated GAC <0.1 Yes [66] Commercial ArsenXnp—a hybrid anion exchanger <1.0 Yes [120] Metal (hydr)oxide Iron mineral-quartz sand 0.4 Yes [16] Mineral Natural siderite 0.4 Yes [45] Mineral Granular schwertmannite 0.025 Yes [46] Mineral Activated siderite-hematite 0.3 Yes [39]

In addition to waste landfill disposal, studies reported other destination alternatives: the incorporation of arsenic sludge in construction material, the mixing of arsenic waste Sustainability 2021, 13, 1186 19 of 25

from biomass origin with coal for energy generation, and the mixing with livestock waste to be converted by microorganisms into gaseous arsine [145].

6. Conclusions Much research has been conducted in the past decades to develop adsorption systems for continuous arsenic removal from contaminated water. Different types of materials have been tested and improved to perform better adsorption. Maximum adsorption capacities of 28.96 mg/g and 74.8 mg/g were reported mostly by novel adsorbents like binary metal oxides and biosorbents, respectively. Most studies were preliminary, and there is still a gap between the identification of novel adsorbents and feasibility studies for practical applications. The other materials presented similar values of maximum adsorption capacities: 6.07 mg/g for GFH, 5.00 mg/g for minerals (mill-scale derived magnetite particles), 3.8852 mg/g for industrial waste (iron oxide-coated cement), 3.51 mg/g for soil and rock (pisolite), 3.30 mg/g for nanocomposite (iron doped phenolic resin/AC), 3.229 mg/g for commercial (Lewatit FO36), and 2.873 mg/g for carbon-based (Fe/GAC), Biosorbents were employed in most studies published, followed by metal oxides and nanocomposites. There was an increase in the last years in research for low-cost adsorbents such as biosorbents and industrial waste and by-products, which is of high importance to fill the gap in efficient and cheap treatment in most arsenic-affected places. Arsenic adsorption in continuous mode is mostly affected by pH, temperature, flow rate and hydraulic retention time, adsorbent quantity (bed height), adsorbate initial concentra- tion, and the presence of other chemical elements. Iron incorporation is the most common treatment to increase the arsenic adsorption capacity of adsorbent materials, followed by other metals such as aluminum and zirconium. Desorption, like adsorption, is dependent on parameters such as eluent type and concentration, pH, flow rate, and retention time. For each arsenic adsorption system, those parameters should be adjusted to result in better regeneration of adsorbent and recovery of adsorbate. Arsenic-contaminated adsorbents can be non-hazardous, and a variety of disposal alternatives are being investigated for the fate of arsenic waste. However, the topic should be the focus of deeper studies, especially in the testing of the viability of exhausted adsor- bents alternative applications on a real scale, including more research into the toxicity and stability of materials incorporating arsenic adsorbent waste. Further adsorption investigation must be conducted at pilot-scale, which is important to evaluate the application of the adsorbent in full scale, including more studies in regen- eration and recovery, which enhance the economical balance of adsorption. Furthermore, the calculation of treatment costs and life cycle assessment (LCA) of the adsorption systems are relevant to be compared with other water treatment techniques. Overall, this work gathered the most relevant worldwide research into low-cost and alternative arsenic adsorption dynamic systems, which is important to help to consolidate the available knowledge for sustainable solutions that provide safe access to potable water in contaminated areas.

Author Contributions: Conceptualization, A.M.A.P.; investigation, M.A.C.; resources, C.M.S.B.; data curation, M.A.C.; writing—original draft preparation, M.A.C.; writing—review and editing, A.M.A.P., C.M.S.B., and R.A.R.B.; visualization, M.A.C.; supervision, A.M.A.P.; funding acquisition, C.M.S.B. and R.A.R.B. All authors have read and agreed to the published version of the manuscript. Funding: This work was financially supported by Base Funding—UIDB/50020/2020 of the Associate Laboratory LSRE-LCM—funded by national funds through FCT/MCTES (PIDDAC). A. Pintor acknowledges her Junior Researcher contract [CEECIND/01485/2017] by FCT. Institutional Review Board Statement: Not applicable. Informed Consent Statement: Not applicable. Sustainability 2021, 13, 1186 20 of 25

Data Availability Statement: Data sharing not applicable. Conflicts of Interest: The authors declare no conflict of interest.

References 1. Villaescusa, I.; Bollinger, J.-C. Arsenic in drinking water: Sources, occurrence and health effects (a review). Rev. Environ. Sci. Bio/Technol. 2008, 7, 307–323. [CrossRef] 2. Smedley, P.L.; Kinniburgh, D.G. A review of the source, behaviour and distribution of arsenic in natural waters. Appl. Geochem. 2002, 17, 517–568. [CrossRef] 3. Ungureanu, G.; Santos, S.; Boaventura, R.; Botelho, C. Arsenic and antimony in water and wastewater: Overview of removal techniques with special reference to latest advances in adsorption. J. Environ. Manag. 2015, 151, 326–342. [CrossRef] 4. Li, Z.; Yang, Q.; Yang, Y.; Xie, C.; Ma, H. Hydrogeochemical controls on arsenic contamination potential and health threat in an intensive agricultural area, northern China. Environ. Pollut. 2020, 256, 113455. [CrossRef] 5. Pintor, A.M.A.; Vieira, B.R.C.; Santos, S.C.R.; Boaventura, R.A.R.; Botelho, C.M.S. Arsenate and arsenite adsorption onto iron-coated cork granulates. Sci. Total Environ. 2018, 642, 1075–1089. [CrossRef] 6. Patel, H. Fixed-bed column adsorption study: A comprehensive review. Appl. Water Sci. 2019, 9, 45. [CrossRef] 7. Zeng, H.; Yu, Y.; Wang, F.; Zhang, J.; Li, D. Arsenic(V) removal by granular adsorbents made from water treatment residuals materials and chitosan. Colloids Surf. Phys. Eng. Asp. 2020, 585, 124036. [CrossRef] 8. Hao, L.; Liu, M.; Wang, N.; Li, G. A critical review on arsenic removal from water using iron-based adsorbents. RSC Adv. 2018, 8, 39545–39560. [CrossRef] 9. Asere, T.G.; Stevens, C.V.; Du Laing, G. Use of (modified) natural adsorbents for arsenic remediation: A review. Sci. Total Environ. 2019, 676, 706–720. [CrossRef] 10. Bhakta, J.N.; Ali, M.M. Biosorption of Arsenic: An Emerging Eco-technology of Arsenic Detoxification in Drinking Water. In Arsenic Water Resources Contamination: Challenges and Solutions; Fares, A., Singh, S.K., Eds.; Springer International Publishing: Cham, switzerland, 2020. [CrossRef] 11. Lata, S.; Samadder, S.R. Removal of arsenic from water using nano adsorbents and challenges: A review. J. Environ. Manag. 2016, 166, 387–406. [CrossRef] 12. Bullen, J.C.; Lapinee, C.; Salaün, P.; Vilar, R.; Weiss, D.J. On the application of photocatalyst-sorbent composite materials for arsenic(III) remediation: Insights from kinetic adsorption modelling. J. Environ. Chem. Eng. 2020, 104033. [CrossRef] 13. Streat, M.; Hellgardt, K.; Newton, N.L.R. Hydrous ferric oxide as an adsorbent in water treatment: Part 1. Preparation and physical characterization. Process Saf. Environ. Prot. 2008, 86, 1–9. [CrossRef] 14. Chang, Y.-Y.; Song, K.-H.; Yang, J.-K. Removal of As(III) in a column reactor packed with iron-coated sand and manganese-coated sand. J. Hazard. Mater. 2008, 150, 565–572. [CrossRef][PubMed] 15. Chaudhry, S.A.; Zaidi, Z.; Siddiqui, S.I. Isotherm, kinetic and thermodynamics of arsenic adsorption onto Iron-Zirconium Binary Oxide-Coated Sand (IZBOCS): Modelling and process optimization. J. Mol. Liq. 2017, 229, 230–240. [CrossRef] 16. Guo, H.; Stüben, D.; Berner, Z. Arsenic removal from water using natural iron mineral–quartz sand columns. Sci. Total Environ. 2007, 377, 142–151. [CrossRef] 17. Khan, A.H.; Rasul, S.B.; Munir, A.K.M.; Habibuddowla, M.; Alauddin, M.; Newaz, S.S.; Hussam, A. Appraisal of a simple arsenic removal method for ground water of Bangladesh. J. Environ. Sci. Health Part A 2000, 35, 1021–1041. [CrossRef] 18. Wang, Y.; Sun, L.; Han, T.; Si, Y.; Wang, R. Arsenite and arsenate leaching and retention on iron (hydr)oxide-coated sand column. J. Soils Sediments 2016, 16, 486–496. [CrossRef] 19. Thirunavukkarasu, O.S.; Viraraghavan, T.; Subramanian, K.S.; Tanjore, S. Organic arsenic removal from drinking water. Urban Water 2002, 4, 415–421. [CrossRef] 20. Manna, B.R.; Dey, S.; Debnath, S.; Ghosh, U.C. Removal of Arsenic from Groundwater using Crystalline Hydrous Ferric Oxide (CHFO). Water Qual. Res. J. 2003, 38, 193–210. [CrossRef] 21. Gupta, K.; Ghosh, U.C. Arsenic removal using hydrous nanostructure iron(III)–titanium(IV) binary mixed oxide from aqueous solution. J. Hazard. Mater. 2009, 161, 884–892. [CrossRef] 22. Chaudhry, S.A.; Khan, T.A.; Ali, I. Zirconium oxide-coated sand based batch and column adsorptive removal of arsenic from water: Isotherm, kinetic and thermodynamic studies. Egypt. J. Pet. 2017, 26, 553–563. [CrossRef] 23. He, X.; Deng, F.; Shen, T.; Yang, L.; Chen, D.; Luo, J.; Luo, X.; Min, X.; Wang, F. Exceptional adsorption of arsenic by zirconium metal-organic frameworks: Engineering exploration and mechanism insight. J. Colloid Interface Sci. 2019, 539, 223–234. [CrossRef] [PubMed] 24. Ouvrard, S.; Simonnot, M.-O.; Sardin, M. Reactive behavior of natural manganese oxides toward the adsorption of phosphate and arsenate. Ind. Eng. Chem. Res. 2002, 41, 2785–2791. [CrossRef] 25. Manna, B.; Ghosh, U.C. Adsorption of arsenic from aqueous solution on synthetic hydrous stannic oxide. J. Hazard. Mater. 2007, 144, 522–531. [CrossRef] 26. Biterna, M.; Arditsoglou, A.; Tsikouras, E.; Voutsa, D. Arsenate removal by zero valent iron: Batch and column tests. J. Hazard. Mater. 2007, 149, 548–552. [CrossRef] 27. Melitas, N.; Wang, J.; Conklin, M.; O’Day, P.; Farrell, J. Understanding Soluble Arsenate Removal Kinetics by Zerovalent Iron Media. Environ. Sci. Technol. 2002, 36, 2074–2081. [CrossRef] Sustainability 2021, 13, 1186 21 of 25

28. Biterna, M.; Antonoglou, L.; Lazou, E.; Voutsa, D. Arsenite removal from waters by zero valent iron: Batch and column tests. Chemosphere 2010, 78, 7–12. [CrossRef] 29. Leupin, O.X.; Hug, S.J. Oxidation and removal of arsenic (III) from aerated groundwater by filtration through sand and zero-valent iron. Water Res. 2005, 39, 1729–1740. [CrossRef] 30. Ludwig, R.D.; Smyth, D.J.A.; Blowes, D.W.; Spink, L.E.; Wilkin, R.T.; Jewett, D.G.; Weisener, C.J. Treatment of Arsenic, Heavy Met- als, and Acidity Using a Mixed ZVI-Compost PRB. Environ. Sci. Technol. 2009, 43, 1970–1976. [CrossRef] 31. Su, C.; Puls, R.W. Significance of Iron(II, III) Hydroxycarbonate Green Rust in Arsenic Remediation Using Zerovalent Iron in Laboratory Column Tests. Environ. Sci. Technol. 2004, 38, 5224–5231. [CrossRef] 32. Deliyanni, E.A.; Bakoyannakis, D.N.; Zouboulis, A.I.; Peleka, E. Removal of Arsenic and Cadmium by Akaganeite Fixed-Beds. Sep. Sci. Technol. 2003, 38, 3967–3981. [CrossRef] 33. Yin, H.; Kong, M.; Gu, X.; Chen, H. Removal of arsenic from water by porous charred granulated attapulgite-supported hydrated iron oxide in bath and column modes. J. Clean. Prod. 2017, 166, 88–97. [CrossRef] 34. Ayoob, S.; Gupta, A.K.; Bhakat, P.B. Analysis of breakthrough developments and modeling of fixed bed adsorption system for As(V) removal from water by modified calcined bauxite (MCB). Sep. Purif. Technol. 2007, 52, 430–438. [CrossRef] 35. Ayoob, S.; Gupta, A.K.; Bhakat, P.B. Performance evaluation of modified calcined bauxite in the sorptive removal of arsenic(III) from aqueous environment. Colloids Surf. Phys. Eng. Asp. 2007, 293, 247–254. [CrossRef] 36. Bhakat, P.B.; Gupta, A.K.; Ayoob, S. Feasibility analysis of As(III) removal in a continuous flow fixed bed system by modified calcined bauxite (MCB). J. Hazard. Mater. 2007, 139, 286–292. [CrossRef] 37. Bhakat, P.B.; Gupta, A.K.; Ayoob, S.; Kundu, S. Investigations on arsenic(V) removal by modified calcined bauxite. Colloids Surf. Phys. Eng. Asp. 2006, 281, 237–245. [CrossRef] 38. Ramirez-Muñiz, K.; Perez-Rodriguez, F.; Rangel-Mendez, R. Adsorption of arsenic onto an environmental friendly goethite- polyacrylamide composite. J. Mol. Liq. 2018, 264, 253–260. [CrossRef] 39. Guo, H.; Stüben, D.; Berner, Z.; Kramar, U. Adsorption of arsenic species from water using activated siderite–hematite column filters. J. Hazard. Mater. 2008, 151, 628–635. [CrossRef] 40. Zhou, J.; Chen, S.; Liu, J.; Frost, R.L. Adsorption kinetic and species variation of arsenic for As(V) removal by biologically mackinawite (FeS). Chem. Eng. J. 2018, 354, 237–244. [CrossRef] 41. Navarathna, C.M.; Karunanayake, A.G.; Gunatilake, S.R.; Pittman, C.U.; Perez, F.; Mohan, D.; Mlsna, T. Removal of Arsenic(III) from water using magnetite precipitated onto Douglas fir biochar. J. Environ. Manag. 2019, 250, 109429. [CrossRef] 42. Shahid, M.K.; Phearom, S.; Choi, Y.-G. Evaluation of arsenate adsorption efficiency of mill-scale derived magnetite particles with column and plug flow reactors. J. Water Process Eng. 2019, 28, 260–268. [CrossRef] 43. Shahid, M.K.; Phearom, S.; Choi, Y.-G. Adsorption of arsenic (V) on magnetite-enriched particles separated from the mill scale. Environ. Earth Sci. 2019, 78, 65. [CrossRef] 44. Shahid, M.K.; Phearom, S.; Choi, Y.-G. Packed Bed Column for Adsorption of Arsenic on Mixed-Valent Iron [Fe(II)-Fe(III)] Oxide Synthesized from Industrial Waste. J. Hazard. Toxic Radioact. Waste 2020, 24, 04020002. [CrossRef] 45. Guo, H.; Stüben, D.; Berner, Z. Adsorption of arsenic(III) and arsenic(V) from groundwater using natural siderite as the adsorbent. J. Colloid Interface Sci. 2007, 315, 47–53. [CrossRef] 46. Dou, X.; Mohan, D.; Pittman, C.U. Arsenate adsorption on three types of granular schwertmannite. Water Res. 2013, 47, 2938–2948. [CrossRef] 47. Baskan, M.B.; Pala, A. Batch and Fixed-Bed Column Studies of Arsenic Adsorption on the Natural and Modified Clinoptilolite. Water Air Soil Pollut. 2013, 225, 1798. [CrossRef] 48. Jeon, C.-S.; Baek, K.; Park, J.-K.; Oh, Y.-K.; Lee, S.-D. Adsorption characteristics of As(V) on iron-coated zeolite. J. Hazard. Mater. 2009, 163, 804–808. [CrossRef] 49. Guo, H.; Stüben, D.; Berner, Z. Removal of arsenic from aqueous solution by natural siderite and hematite. Appl. Geochem. 2007, 22, 1039–1051. [CrossRef] 50. Maji, S.K.; Pal, A.; Pal, T. Arsenic removal from real-life groundwater by adsorption on laterite soil. J. Hazard. Mater. 2008, 151, 811–820. [CrossRef] 51. Maji, S.K.; Pal, A.; Pal, T.; Adak, A. Modeling and fixed bed column adsorption of As(III) on laterite soil. Sep. Purif. Technol. 2007, 56, 284–290. [CrossRef] 52. Maji, S.K.; Pal, A.; Pal, T.; Adak, A. Modeling and fixed bed column adsorption of As(V) on laterite soil. J. Environ. Sci. Health Part A 2007, 42, 1585–1593. [CrossRef][PubMed] 53. Maiti, A.; DasGupta, S.; Basu, J.K.; De, S. Batch and Column Study: Adsorption of Arsenate Using Untreated Laterite as Adsorbent. Ind. Eng. Chem. Res. 2008, 47, 1620–1629. [CrossRef] 54. Mondal, R.; Mondal, S.; Kurada, K.V.; Bhattacharjee, S.; Sengupta, S.; Mondal, M.; Karmakar, S.; De, S.; Griffiths, I.M. Modelling the transport and adsorption dynamics of arsenic in a soil bed filter. Chem. Eng. Sci. 2019, 210, 115205. [CrossRef] 55. Williams, L.E.; Barnett, M.O.; Kramer, T.A.; Melville, J.G. Adsorption and Transport of Arsenic(V) in Experimental Subsurface Systems. J. Environ. Qual. 2003, 32, 841–850. [CrossRef][PubMed] 56. Pereira, P.A.L.; Dutra, A.J.B.; Martins, A.H. Adsorptive removal of arsenic from river waters using pisolite. Miner. Eng. 2007, 20, 52–59. [CrossRef] Sustainability 2021, 13, 1186 22 of 25

57. Sosa, A.; Armienta, M.A.; Aguayo, A.; Cruz, O. Evaluation of the influence of main groundwater ions on arsenic removal by limestones through column experiments. Sci. Total Environ. 2020, 727, 138459. [CrossRef][PubMed] 58. Jang, M.; Min, S.H.; Kim, T.H.; Park, J.K. Removal of arsenite and arsenate using hydrous ferric oxide incorporated into naturally occurring porous diatomite. Environ. Sci. Technol. 2006, 40, 1636–1643. [CrossRef] 59. Maji, S.K.; Kao, Y.-H.; Wang, C.-J.; Lu, G.-S.; Wu, J.-J.; Liu, C.-W. Fixed bed adsorption of As(III) on iron-oxide-coated natural rock (IOCNR) and application to real arsenic-bearing groundwater. Chem. Eng. J. 2012, 203, 285–293. [CrossRef] 60. Maji, S.K.; Kao, Y.-H.; Wang, Y.-B.; Liu, C.-W. Dynamic column adsorption of As on iron-oxide-coated natural rock (IOCNR) and sludge management. Desalination Water Treat. 2015, 55, 2171–2182. [CrossRef] 61. Chen, W.; Parette, R.; Zou, J.; Cannon, F.S.; Dempsey, B.A. Arsenic removal by iron-modified activated carbon. Water Res. 2007, 41, 1851–1858. [CrossRef] 62. Chen, R.; Zhang, Z.; Lei, Z.; Sugiura, N. Preparation of iron-impregnated tablet ceramic adsorbent for arsenate removal from aqueous solutions. Desalination 2012, 286, 56–62. [CrossRef] 63. Kalaruban, M.; Loganathan, P.; Nguyen, T.V.; Nur, T.; Hasan Johir, M.A.; Nguyen, T.H.; Trinh, M.V.; Vigneswaran, S. Iron-impregnated granular activated carbon for arsenic removal: Application to practical column filters. J. Environ. Manag. 2019, 239, 235–243. [CrossRef][PubMed] 64. Sigrist, M.E.; Brusa, L.; Beldomenico, H.R.; Dosso, L.; Tsendra, O.M.; González, M.B.; Pieck, C.L.; Vera, C.R. Influence of the iron content on the arsenic adsorption capacity of Fe/GAC adsorbents. J. Environ. Chem. Eng. 2014, 2, 927–934. [CrossRef] 65. Zhang, Q.L.; Lin, Y.C.; Chen, X.; Gao, N.Y. A method for preparing ferric activated carbon composites adsorbents to remove arsenic from drinking water. J. Hazard. Mater. 2007, 148, 671–678. [CrossRef][PubMed] 66. Zhuang, J.M.; Hobenshield, E.; Walsh, T. Arsenate sorption by hydrous ferric oxide incorporated onto granular activated carbon with phenol formaldehyde resins coating. Environ. Technol. 2008, 29, 401–411. [CrossRef] 67. Gu, Z.; Fang, J.; Deng, B. Preparation and Evaluation of GAC-Based Iron-Containing Adsorbents for Arsenic Removal. Environ. Sci. Technol. 2005, 39, 3833–3843. [CrossRef] 68. Igwe, J.; Abia, A. A bioseparation process for removing heavy metals from waste water using biosorbents. Afr. J. Biotechnol. 2006, 5, 1167–1179. [CrossRef] 69. Saqib, A.N.S.; Waseem, A.; Khan, A.F.; Mahmood, Q.; Khan, A.; Habib, A.; Khan, A.R. Arsenic bioremediation by low cost materials derived from Blue Pine (Pinus wallichiana) and Walnut (Juglans regia). Ecol. Eng. 2013, 51, 88–94. [CrossRef] 70. Boddu, V.M.; Abburi, K.; Talbott, J.L.; Smith, E.D.; Haasch, R. Removal of arsenic (III) and arsenic (V) from aqueous medium using chitosan-coated biosorbent. Water Res. 2008, 42, 633–642. [CrossRef] 71. Brion-Roby, R.; Gagnon, J.; Deschênes, J.-S.; Chabot, B. Investigation of fixed bed adsorption column operation parameters using a chitosan material for treatment of arsenate contaminated water. J. Environ. Chem. Eng. 2018, 6, 505–511. [CrossRef] 72. Chen, C.-C.; Chung, Y.-C. Arsenic Removal Using a Biopolymer Chitosan Sorbent. J. Environ. Sci. Health Part A 2006, 41, 645–658. [CrossRef][PubMed] 73. Chen, C.-Y.; Chang, T.-H.; Kuo, J.-T.; Chen, Y.-F.; Chung, Y.-C. Characteristics of molybdate-impregnated chitosan beads (MICB) in terms of arsenic removal from water and the application of a MICB-packed column to remove arsenic from wastewater. Bioresour. Technol. 2008, 99, 7487–7494. [CrossRef][PubMed] 74. Dhoble, R.M.; Maddigapu, P.R.; Rayalu, S.S.; Bhole, A.G.; Dhoble, A.S.; Dhoble, S.R. Removal of arsenic(III) from water by magnetic binary oxide particles (MBOP): Experimental studies on fixed bed column. J. Hazard. Mater. 2017, 322, 469–478. [CrossRef][PubMed] 75. Gupta, A.; Chauhan, V.S.; Sankararamakrishnan, N. Preparation and evaluation of iron–chitosan composites for removal of As(III) and As(V) from arsenic contaminated real life groundwater. Water Res. 2009, 43, 3862–3870. [CrossRef] 76. Gupta, A.; Sankararamakrishnan, N. Column studies on the evaluation of novel spacer granules for the removal of arsenite and arsenate from contaminated water. Bioresour. Technol. 2010, 101, 2173–2179. [CrossRef] 77. Kang, S.; Park, S.-M.; Park, J.-G.; Baek, K. Enhanced adsorption of arsenic using calcined alginate bead containing alum sludge from water treatment facilities. J. Environ. Manag. 2019, 234, 181–188. [CrossRef] 78. Oci´nski,D.; Mazur, P. Highly efficient arsenic sorbent based on residual from water deironing—Sorption mechanisms and column studies. J. Hazard. Mater. 2020, 382, 121062. [CrossRef] 79. Rahaman, M.S.; Basu, A.; Islam, M.R. The removal of As(III) and As(V) from aqueous solutions by waste materials. Bioresour. Technol. 2008, 99, 2815–2823. [CrossRef] 80. Roxanne, B.-R.; Jonathan, G.; Jean-Sébastien, D.; Bruno, C. Development and treatment procedure of arsenic-contaminated water using a new and green chitosan sorbent: Kinetic, isotherm, thermodynamic and dynamic studies. Pure Appl. Chem. 2018, 90, 63–77. [CrossRef] 81. Saha, S.; Sarkar, P. Arsenic mitigation by chitosan based porous magnesia impregnated alumina: Performance evaluation in continuous packed bed column. Int. J. Environ. Sci. Technol. 2015, 13.[CrossRef] 82. Sarkar, P.; Pal, P.; Bhattacharyay, D.; Banerjee, S. Removal of arsenic from drinking water by ferric hydroxide microcapsule-loaded alginate beads in packed adsorption column. J. Environ. Sci. Health Part A 2010, 45, 1750–1757. [CrossRef][PubMed] 83. Wei, Y.; Yu, X.; Liu, C.; Ma, J.; Wei, S.; Chen, T.; Yin, K.; Liu, H.; Luo, S. Enhanced arsenite removal from water by radially porous Fe-chitosan beads: Adsorption and H2O2 catalytic oxidation. J. Hazard. Mater. 2019, 373, 97–105. [CrossRef][PubMed] Sustainability 2021, 13, 1186 23 of 25

84. Zouboulis, A.I.; Katsoyiannis, I.A. Arsenic Removal Using Iron Oxide Loaded Alginate Beads. Ind. Eng. Chem. Res. 2002, 41, 6149–6155. [CrossRef] 85. Ghimire, K.N.; Inoue, K.; Makino, K.; Miyajima, T. Adsorptive removal of arsenic using orange juice residue. Sep. Sci. Technol. 2002, 37, 2785–2799. [CrossRef] 86. Haque, M.N.; Morrison, G.M.; Perrusquía, G.; Gutierréz, M.; Aguilera, A.F.; Cano-Aguilera, I.; Gardea-Torresdey, J.L. Characteris- tics of arsenic adsorption to sorghum biomass. J. Hazard. Mater. 2007, 145, 30–35. [CrossRef] 87. Hoshina, H.; Takahashi, M.; Kasai, N.; Seko, N. Adsorbent for Arsenic(V) Removal Synthesized by Radiation-Induced Graft Polymerization onto Nonwoven Cotton Fabric. Int. J. Org. Chem. 2012, 2, 173–177. [CrossRef] 88. Hu, X.; Ding, Z.; Zimmerman, A.R.; Wang, S.; Gao, B. Batch and column sorption of arsenic onto iron-impregnated biochar synthesized through hydrolysis. Water Res. 2015, 68, 206–216. [CrossRef] 89. Jaiswal, V.; Saxena, S.; Kaur, I.; Dubey, P.; Nand, S.; Naseem, M.; Singh, S.B.; Srivastava, P.K.; Barik, S.K. Application of four novel fungal strains to remove arsenic from contaminated water in batch and column modes. J. Hazard. Mater. 2018, 356, 98–107. [CrossRef] 90. Kamala, C.T.; Chu, K.H.; Chary, N.S.; Pandey, P.K.; Ramesh, S.L.; Sastry, A.R.; Sekhar, K.C. Removal of arsenic(III) from aqueous solutions using fresh and immobilized plant biomass. Water Res. 2005, 39, 2815–2826. [CrossRef] 91. Pang, D.; Wang, C.-C.; Wang, P.; Liu, W.; Fu, H.; Zhao, C. Superior removal of inorganic and organic arsenic pollutants from water with MIL-88A(Fe) decorated on cotton fibers. Chemosphere 2020, 254, 126829. [CrossRef] 92. Pokhrel, D.; Viraraghavan, T. Arsenic removal in an iron oxide-coated fungal biomass column: Analysis of breakthrough curves. Bioresour. Technol. 2008, 99, 2067–2071. [CrossRef][PubMed] 93. Ranjan, D.; Talat, M.; Hasan, S.H. Rice Polish: An Alternative to Conventional Adsorbents for Treating Arsenic Bearing Water by Up-Flow Column Method. Ind. Eng. Chem. Res. 2009, 48, 10180–10185. [CrossRef] 94. Roy, P.; Mondal, N.K.; Bhattacharya, S.; Das, B.; Das, K. Removal of arsenic(III) and arsenic(V) on chemically modified low-cost adsorbent: Batch and column operations. Appl. Water Sci. 2013, 3, 293–309. [CrossRef] 95. Samad, A.; Fukumoto, T.; Dabwan, A.; Katsumata, H.; Suzuki, T.; Furukawa, M.; Knaeco, S. Enhanced Removal of Arsenite from Ground Water by Adsorption onto Heat-Treated Rice Husk. Open J. Inorg. Non-Met. Mater. 2016, 6, 18–23. [CrossRef] 96. Wei, Y.; Wei, S.; Liu, C.; Chen, T.; Tang, Y.; Ma, J.; Yin, K.; Luo, S. Efficient removal of arsenic from groundwater using iron oxide nanoneedle array-decorated biochar fibers with high Fe utilization and fast adsorption kinetics. Water Res. 2019, 167, 115107. [CrossRef] 97. Zhao, Y.; Huang, M.; Wu, W.; Jin, W. Synthesis of the cotton cellulose based Fe(III)-loaded adsorbent for arsenic(V) removal from drinking water. Desalination 2009, 249, 1006–1011. [CrossRef] 98. Liu, Y.; Chen, S.; Zhong, L.; Wu, G. Preparation of high-stable silver nanoparticle dispersion by using sodium alginate as a stabilizer under gamma radiation. Radiat. Phys. Chem. 2009, 78, 251–255. [CrossRef] 99. Hassan, A.F.; Abdel-Mohsen, A.M.; Elhadidy, H. Adsorption of arsenic by activated carbon, calcium alginate and their composite beads. Int. J. Biol. Macromol. 2014, 68, 125–130. [CrossRef] 100. Yuan, T.; Yong Hu, J.; Ong, S.L.; Luo, Q.F.; Jun Ng, W. Arsenic removal from household drinking water by adsorption. J. Environ. Sci. Health Part A 2002, 37, 1721–1736. [CrossRef] 101. Singh, P.; Singh, S.K.; Bajpai, J.; Bajpai, A.K.; Shrivastava, R.B. Iron crosslinked alginate as novel nanosorbents for removal of arsenic ions and bacteriological contamination from water. J. Mater. Res. Technol. 2014, 3, 195–202. [CrossRef] 102. Mustafiz, S.; Basu, A.; Islam, M.R.; Dewaidar, A.; Chaalal, O. A Novel Method for Heavy Metal Removal. Energy Sources 2002, 24, 1043–1051. [CrossRef] 103. Basu, A.; Mustafiz, S.; Islam, M.R.; Bjorndalen, N.; Rahaman, M.S.; Chaalal, O. A comprehensive approach for modeling sorption of lead and ions through fish scales as an adsorbent. Chem. Eng. Commun. 2006, 193, 580–605. [CrossRef] 104. Sud, D.; Mahajan, G.; Kaur, M.P. Agricultural waste material as potential adsorbent for sequestering heavy metal ions from aqueous solutions—A review. Bioresour. Technol. 2008, 99, 6017–6027. [CrossRef][PubMed] 105. Jung, K.-W.; Kim, K.; Jeong, T.-U.; Ahn, K.-H. Influence of pyrolysis temperature on characteristics and phosphate adsorption capability of biochar derived from waste-marine macroalgae (Undaria pinnatifida roots). Bioresour. Technol. 2016, 200, 1024–1028. [CrossRef][PubMed] 106. Tan, X.; Liu, Y.; Zeng, G.; Wang, X.; Hu, X.; Gu, Y.; Yang, Z. Application of biochar for the removal of pollutants from aqueous solutions. Chemosphere 2015, 125, 70–85. [CrossRef] 107. Ali, I.; Al-Othman, Z.A.; Alwarthan, A.; Asim, M.; Khan, T.A. Removal of arsenic species from water by batch and column operations on bagasse fly ash. Environ. Sci. Pollut. Res. 2014, 21, 3218–3229. [CrossRef] 108. Gibbons, M.K.; Gagnon, G.A. Adsorption of arsenic from a Nova Scotia groundwater onto water treatment residual solids. Water Res. 2010, 44, 5740–5749. [CrossRef] 109. Kundu, S.; Gupta, A.K. Analysis and modeling of fixed bed column operations on As(V) removal by adsorption onto iron oxide-coated cement (IOCC). J. Colloid Interface Sci. 2005, 290, 52–60. [CrossRef] 110. Kundu, S.; Gupta, A.K. As(III) removal from aqueous medium in fixed bed using iron oxide-coated cement (IOCC): Experimental and modeling studies. Chem. Eng. J. 2007, 129, 123–131. [CrossRef] 111. Kundu, S.; Kavalakatt, S.S.; Pal, A.; Ghosh, S.K.; Mandal, M.; Pal, T. Removal of arsenic using hardened paste of Portland cement: Batch adsorption and column study. Water Res. 2004, 38, 3780–3790. [CrossRef] Sustainability 2021, 13, 1186 24 of 25

112. Liem-Nguyen, V.; Sjöberg, V.; Dinh, N.P.; Huy, D.H.; Karlsson, S. Removal mechanism of arsenic (V) by stainless steel slags obtained from scrap metal recycling. J. Environ. Chem. Eng. 2020, 8, 103833. [CrossRef] 113. Christian, P.; Von der Kammer, F.; Baalousha, M.; Hofmann, T. Nanoparticles: Structure, properties, preparation and behaviour in environmental media. Ecotoxicology 2008, 17, 326–343. [CrossRef][PubMed] 114. Önnby, L.; Pakade, V.; Mattiasson, B.; Kirsebom, H. Polymer composite adsorbents using particles of molecularly imprinted polymers or aluminium oxide nanoparticles for treatment of arsenic contaminated waters. Water Res. 2012, 46, 4111–4120. [CrossRef][PubMed] 115. Ali, I. Microwave assisted economic synthesis of multi walled carbon nanotubes for arsenic species removal in water: Batch and column operations. J. Mol. Liq. 2018, 271, 677–685. [CrossRef] 116. Sharma, A.; Verma, N.; Sharma, A.; Deva, D.; Sankararamakrishnan, N. Iron doped phenolic resin based activated carbon micro and nanoparticles by milling: Synthesis, characterization and application in arsenic removal. Chem. Eng. Sci. 2010, 65, 3591–3601. [CrossRef] 117. Wang, J.; Zhang, W.; Zheng, Y.; Zhang, N.; Zhang, C. Multi-functionalization of magnetic graphene by surface-initiated ICAR ATRP mediated by polydopamine chemistry for adsorption and speciation of arsenic. Appl. Surf. Sci. 2019, 478, 15–25. [CrossRef] 118. Contreras, S.; Henríquez-Vargas, L.; Álvarez, P.I. Arsenic transport and adsorption modeling in columns using a copper nanoparticles composite. J. Water Process Eng. 2017, 19, 212–219. [CrossRef] 119. Reddy, K.J.; McDonald, K.J.; King, H. A novel arsenic removal process for water using cupric oxide nanoparticles. J. Colloid Interface Sci. 2013, 397, 96–102. [CrossRef] 120. Sarkar, S.; Blaney, L.M.; Gupta, A.; Ghosh, D.; SenGupta, A.K. Use of ArsenXnp, a hybrid anion exchanger, for arsenic removal in remote villages in the Indian subcontinent. React. Funct. Polym. 2007, 67, 1599–1611. [CrossRef] 121. Morillo, D.; Faccini, M.; Amantia, D.; Pérez, G.; García, M.A.; Valiente, M.; Aubouy, L. Superparamagnetic iron oxide nanoparticle- loaded polyacrylonitrile nanofibers with enhanced arsenate removal performance. Environ. Sci. Nano 2016, 3, 1165–1173. [CrossRef] 122. Hernández-Flores, H.; Pariona, N.; Herrera-Trejo, M.; Hdz-García, H.M.; Mtz-Enriquez, A.I. Concrete/maghemite nanocomposites as novel adsorbents for arsenic removal. J. Mol. Struct. 2018, 1171, 9–16. [CrossRef] 123. Luo, J.; Meng, X.; Crittenden, J.; Qu, J.; Hu, C.; Liu, H.; Peng, P. Arsenic adsorption on α-MnO2 nanofibers and the significance of (100) facet as compared with (110). Chem. Eng. J. 2018, 331, 492–500. [CrossRef] 124. Chu, K.H. Prediction of arsenic breakthrough in a pilot column of polymer-supported nanoparticles. J. Water Process Eng. 2014, 3, 117–122. [CrossRef] 125. Savina, I.N.; English, C.J.; Whitby, R.L.D.; Zheng, Y.; Leistner, A.; Mikhalovsky, S.V.; Cundy, A.B. High efficiency removal of dissolved As(III) using iron nanoparticle-embedded macroporous polymer composites. J. Hazard. Mater. 2011, 192, 1002–1008. [CrossRef][PubMed] 126. Pandi, K.; Lee, D.-W.; Choi, J. Facile synthesis of zirconium-organic frameworks@biomass-derived porous graphitic nanocompos- ites: Arsenic adsorption performance and mechanism. J. Mol. Liq. 2020, 314, 113552. [CrossRef] 127. Danish, M.I.; Qazi, I.A.; Zeb, A.; Habib, A.; Awan, M.A.; Khan, Z. Arsenic Removal from Aqueous Solution Using Pure and Metal-Doped Titania Nanoparticles Coated on Glass Beads: Adsorption and Column Studies. J. Nanomater. 2013, 2013, 873694. [CrossRef] 128. Hristovski, K.; Baumgardner, A.; Westerhoff, P. Selecting metal oxide nanomaterials for arsenic removal in fixed bed columns: From nanopowders to aggregated nanoparticle media. J. Hazard. Mater. 2007, 147, 265–274. [CrossRef] 129. Saldaña-Robles, A.; Damian-Ascencio, C.E.; Guerra-Sanchez, R.J.; Saldaña-Robles, A.L.; Saldaña-Robles, N.; Gallegos-Muñoz, A.; Cano-Andrade, S. Effects of the presence of organic matter on the removal of arsenic from groundwater. J. Clean. Prod. 2018, 183, 720–728. [CrossRef] 130. Streat, M.; Hellgardt, K.; Newton, N.L.R. Hydrous ferric oxide as an adsorbent in water treatment: Part 3: Batch and mini-column adsorption of arsenic, , fluorine and cadmium ions. Process Saf. Environ. Prot. 2008, 21–30. [CrossRef] 131. Genç-Fuhrman, H.; Wu, P.; Zhou, Y.; Ledin, A. Removal of As, Cd, Cr, Cu, Ni and Zn from polluted water using an iron based sorbent. Desalination 2008, 226, 357–370. [CrossRef] 132. Boldaji, M.R.; Nabizadeh, R.; Dehghani, M.H.; Nadafi, K.; Mahvi, A.H. Evaluating the performance of iron nanoparticle resin in removing arsenate from water. J. Environ. Sci. Health Part A 2010, 45, 946–950. [CrossRef][PubMed] 133. Ciopec, M.; Negrea, A.; Lupa, L.; Davidescu, C.M.; Negrea, P. Studies Regarding As(V) Adsorption from Underground Water by Fe-XAD8-DEHPA Impregnated Resin. Equilibrium Sorption and Fixed-Bed Column Tests. Molecules 2014, 19, 16082–16101. [CrossRef][PubMed] 134. Dabizha, A.; Bahr, C.; Kersten, M. Predicting breakthrough of vanadium in fixed-bed absorbent columns with complex groundwa- ter chemistries: A multi-component granular ferric hydroxide−vanadate−arsenate−phosphate−silicic acid system. Water Res. X 2020, 9, 100061. [CrossRef][PubMed] 135. Badruzzaman, M.; Westerhoff, P.; Knappe, D.R.U. Intraparticle diffusion and adsorption of arsenate onto granular ferric hydroxide (GFH). Water Res. 2004, 38, 4002–4012. [CrossRef] 136. Fleming, D.E.B.; Eddy, I.S.; Gherase, M.R.; Gibbons, M.K.; Gagnon, G.A. Real-time monitoring of arsenic filtration by granular ferric hydroxide. Appl. Radiat. Isot. 2010, 68, 821–824. [CrossRef] Sustainability 2021, 13, 1186 25 of 25

137. Nguyen, V.L.; Chen, W.-H.; Young, T.; Darby, J. Effect of interferences on the breakthrough of arsenic: Rapid small scale column tests. Water Res. 2011, 45, 4069–4080. [CrossRef] 138. Sperlich, A.; Schimmelpfennig, S.; Baumgarten, B.; Genz, A.; Amy, G.; Worch, E.; Jekel, M. Predicting anion breakthrough in granular ferric hydroxide (GFH) adsorption filters. Water Res. 2008, 42, 2073–2082. [CrossRef] 139. Sperlich, A.; Werner, A.; Genz, A.; Amy, G.; Worch, E.; Jekel, M. Breakthrough behavior of granular ferric hydroxide (GFH) fixed-bed adsorption filters: Modeling and experimental approaches. Water Res. 2005, 39, 1190–1198. [CrossRef] 140. Streat, M.; Hellgardt, K.; Newton, N.L.R. Hydrous ferric oxide as an adsorbent in water treatment: Part 2. Adsorption studies. Process Saf. Environ. Prot. 2008, 86, 11–20. [CrossRef] 141. Dias, A.C.; Fontes, M.P.F. Arsenic (V) removal from water using hydrotalcites as adsorbents: A critical review. Appl. Clay Sci. 2020, 191, 105615. [CrossRef] 142. Nhat Ha, H.N.; Kim Phuong, N.T.; Boi An, T.; Mai Tho, N.T.; Ngoc Thang, T.; Quang Minh, B.; Van Du, C. Arsenate removal by layered double hydroxides embedded into spherical polymer beads: Batch and column studies. J. Environ. Sci. Health Part A 2016, 51, 403–413. [CrossRef][PubMed] 143. Lu, H.; Zhu, Z.; Zhang, H.; Zhu, J.; Qiu, Y. Simultaneous removal of arsenate and antimonate in simulated and practical water samples by adsorption onto Zn/Fe layered double hydroxide. Chem. Eng. J. 2015, 276, 365–375. [CrossRef] 144. Dadwhal, M.; Ostwal, M.M.; Liu, P.K.T.; Sahimi, M.; Tsotsis, T.T. Adsorption of Arsenic on Conditioned Layered Double Hydroxides: Column Experiments and Modeling. Ind. Eng. Chem. Res. 2009, 48, 2076–2084. [CrossRef] 145. Mondal, M.K.; Garg, R. A comprehensive review on removal of arsenic using activated carbon prepared from easily available waste materials. Environ. Sci. Pollut. Res. 2017, 24, 13295–13306. [CrossRef][PubMed]